You are on page 1of 12

Chemical Engineering & Processing: Process Intensification 189 (2023) 109394

Contents lists available at ScienceDirect

Chemical Engineering and Processing - Process


Intensification
journal homepage: www.elsevier.com/locate/cep

Development of a disposable and easy-to-fabricate microfluidic PCR device


for DNA amplification
Hirad Mashouf a, Bahram Talebjedi b, Nishat Tasnim a, Maia Tan b, Sahar Alousi b,
Sepideh Pakpour b, Mina Hoorfar a, *
a
Department of Mechanical Engineering, University of Victoria, Victoria, Canada
b
School of Engineering, University of British Columbia Okanagan Campus, Kelowna, Canada

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

Keywords: The production of microfluidic PCR devices within lab settings in a straightforward process and large numbers is
Microfluidic PCR limited due to complex methods of fabrication and high cost of development. In this study, we developed a
Lambda (λ) DNA amplification unique disposable, low-cost, and easy-to-fabricate microfluidic PCR chip, comprised of a passive micromixer,
Simple fabrication
reaction well, microheater, and a dynamic temperature control system. In addition to being compact, this device
Low-cost PCR chip
has the advantages of microfluidic PCR systems, such as low power and sample consumption, as well as being
disposable microfluidic device
disposable and easy to fabricate. The disposable part of the chip is easy to fabricate and eliminates the run-to-run
carryover contamination while its integration with a reusable heating system reduces the cost of producing a
micro PCR device. The geometry of the micromixer and the pattern of microheater electrodes are analyzed
through experiments to achieve comparable amplification to a conventional thermocycler. The temperature
control system is designed to monitor the temperature within the reaction well with ±1 ◦ C precision. Finally, the
reliability and repeatability of the microfluidic device were verified by amplifying lambda (λ) DNA. The pro­
posed device demonstrated high-quality amplification comparable to the conventional PCR and is an adaptable
and low-cost on-chip PCR device for reliable molecular diagnostics.

1. Introduction introduced to make use of microfluidics in the fields of biomedicine and


chemistry. With microfluidics PCR devices, PCR reactions can be per­
Infectious disease detection is quite challenging in the early stages formed with much smaller reaction volumes, saving reagent costs and
since the pathogenic genetic material is present at low concentrations in reducing waste [12,13]. As a result of their smaller reaction volumes and
biological samples. Polymerase chain reaction (PCR) based molecular their ability to heat and cool small volumes of liquid rapidly, micro­
diagnostics enable fast and accurate detection of infectious disease fluidic PCR devices can achieve faster reaction times. It is possible to
agents [1–3]. To carry out the PCR reaction, the mixture undergoes improve the efficiency of PCR reactions with microfluidics devices by
thermocycling process consisting of periodic heating and cooling be­ controlling the reaction conditions more precisely and by minimizing
tween constant setpoints specified by a particular PCR protocol. Ther­ the effects of thermal gradients and diffusion. One of the main advan­
mocyclers control the dwell time and temperature of each step according tages of the suggested device over previously reported ones is its capa­
to the protocol. To develop accurate point-of-care PCR-based molecular bility in large scale production without needing the industrial
diagnostics, it is necessary to manufacture compact, low-cost, and techniques and equipment. As an attempt in production of micro PCR
easy-to-operate thermocyclers [4,5]. devices with a low power consumption and resource settings, Karprou
Conventional thermal cyclers are bulky, expensive, energy-intensive et al. [14] introduced a fast, continuous-flow μPCR device with inte­
and require sample preparation [6,7]. Microfluidics PCR devices offer grated resistive microheaters. The device was fabricated using
several significant advantages over traditional PCR devices including a PCB-compatible, industrial materials and processes, all on the same PCB
high surface to volume ratio, reduction in the quantity of reactants substrate. In another study, Moschou et al. [15], introduced a DNA
consumed, and high controllability [8–11]. The development of micro­ amplification featuring the integrated heating elements and was fabri­
fluidic PCR equipment began since the idea of a lab-on-a-chip was cated on a commercially available thin polymeric substrate (Pyralux®

* Corresponding author.

https://doi.org/10.1016/j.cep.2023.109394
Received 28 January 2023; Received in revised form 4 April 2023; Accepted 21 April 2023
Available online 22 April 2023
0255-2701/© 2023 Elsevier B.V. All rights reserved.
H. Mashouf et al. Chemical Engineering and Processing - Process Intensification 189 (2023) 109394

Polyimide). The suggested chip had a low power consumption and fast micro heater unit. The designed temperature control system with NTC
amplification rates due to the small thermal mass of the chip and the low thermistor in this study can perform PCR at any temperature inside the
thermal diffusivity of the polymeric substrate where the heating ele­ ordinary working range of PCR protocols. To further reduce energy re­
ments are located. The proposed device can be produced in lab settings quirements, the design employs a passive micromixer with channel ge­
in large numbers without complex fabrication procedures. ometry that relies on molecular diffusion to efficiently mix PCR
Microfluidic PCR devices can be categorized into two groups: space reagents. To prevent washing the reaction chamber after each use or
domain and time domain PCR devices [16,17]. In space domain PCR, the disposing the chip as whole, our developed microfluidic device is
sample is circulated through various predetermined temperature zones composed of a reusable and a disposable section. The disposable section
[18–20]. Conversely, samples are injected and kept still in a reaction is replaced with a new one upon completion of each PCR reaction. All
chamber while a heating system cycles the temperature in time domain components are integrated in a layered design with the aim to reduce the
PCR devices [21–25]. Controlling the sample in space domain PCR is fabrication cost and eliminate the chance of sample-to-sample cross-
challenging due to the constant movement of the sample, requiring contamination. The temperature control unit is programmed to main­
frequent protocol adjustments. In contrast, time domain PCR allows for tain temperature with less than 1 ◦ C error inside common working range
straightforward adjustments to dwelling time and temperature set­ of PCR protocols. The performance of the microfluidic device is verified
points. Furthermore, conducting PCR using different volumes of sample by amplifying lambda DNA (λ-DNA) samples.
is much simpler to achieve in time domain devices. Additionally, space
domain PCR requires more equipment and complex channel designs, 2. Methodology
making it more expensive and challenging to implement. Due to its
multilayered design, our device is simple and low-cost, requiring no 2.1. Working principle
complex channel designs or microvalves. If the device were to be clas­
sified as a space domain PCR, it would likely require microvalves. The proposed compact and easy-to-fabricate microfluidic PCR device
Nevertheless, there are still concerns about the cost and complexity of is developed by integrating several components to create separate
the fabrication process that make these chips inaccessible for large-scale disposable and reusable elements. Fig. 1b illustrates an exploded view of
PCR testing [26–29]. the microchip comprising a micromixer, reaction chamber well, cover­
Although a wide variety of highly sophisticated chips are increas­ slip, microheater, and the temperature control system. Thin film resistor
ingly being demonstrated, few are or can be adopted by the market and acting as the microheater electrode is patterned on a glass substrate
reach the end-users, which could revolutionize healthcare, biology, and while the temperature sensor is attached to backside of it forming the
chemistry. Commercial manufacturing technology lacks a strong foun­ reusable part of the device. Polydimethylsiloxane (PDMS) is the material
dation for large-scale production [30]. To develop low-cost and used for creating micromixer and the reaction well. PDMS is a well-
low-power microfluidic PCR devices, the temperature control unit of known biocompatible material for biological applications and has
must have low fabrication cost and complexity. Previous studies have been commonly used for the fabrication of reaction chambers. It has
reported different types of external heat sources such as good light permeability, adheres firmly to flat surfaces, and is suitable
temperature-controlled blocks [31], Peltier elements [32], and infrared for mass production [47,48]. These two components are fabricated via
lamps [33]. These methods have considerable heat capacity, high power soft photolithography and then bonded onto a coverslip.
consumption, increase the size of the device and need bulky and Assembling the disposable part on a coverslip eliminates sample-to-
expensive external components, which are not suitable for developing a sample contamination while channels and reaction chambers in micro­
low-cost and compact microfluidic PCR device. Nonetheless, thin film fluidic devices, which are non-disposable are often washed upon PCR
resistance heaters achieve fast heating/cooling rates and effective power completion. Therefore, reagents or washing liquids may be left in
consumption because they have low thermal masses and can be fabri­ channels or absorbed by PDMS due to their porous structure leading to
cated near the reaction chamber. Platinum [34–36], gold [37,38], distortion in the next amplification process results. In this study, it is
Aluminum [39], indium tin oxide (ITO) [40], and several thin film proposed to bond the disposable part on a coverslip. If the disposable
heaters have been patterned on microfluidic chips for PCR applications. part made of PDMS were bonded directly to the glass slide (on which
Thin film Platinum has been widely used as a heater due to its high electrodes are deposited), the whole device would have to be discarded
temperature coefficient of resistivity and linear behavior, while the low after each test since it is not possible to separate the PDMS block from
resistance of gold makes it most suitable to be utilized as electrical leads the glass slide. Even after separation, the electrodes would be impacted
[41]. In microfluidic chips, it is possible to combine the heater and and unlikely to function. There must be a biologically clean layer to
sensor into a single component. Linear behavior and high sensitivity are insulate the PCR mixture from contacting electrodes directly. The elec­
desired characteristics in temperature sensors. Platinum’s stability and trodes could be coated with SiO2 and then bonded with PDMS, but that
high resistance allows it to be simultaneously used as thin film for would increase the fabrication cost.
heating and temperature sensing in microfluidic devices. In these chips, The advantage of having a separate PDMS block for reaction well and
sensors are usually narrow and small, with higher resistance compared bonding the micromixer block on top of it is cutting out the need for a
to the thin film heater to have a fast thermal response [41–43]. Never­ microvalve. Integrating microvalves into PCR microfluidic devices
theless, platinum is expensive, and its complicated and dangerous makes the fabrication process much more challenging, and supple­
fabrication procedure involving the lift-off process [44] and wet etching mentary equipment, such as an air compressor, is required for the valves
in toxic aqua regia [45] complicates its incorporation in microfluidic to operate leading to increase in size and cost of the device. Further­
PCR devices. In a study conducted by Kaprou et al. [46], the authors more, since microvalves do not always perform ideally, some of the PCR
introduced a static μPCR devices with resistive microheaters integrated mixture might flow back to the channels during the thermocycling
on PCBs as miniaturized thermocyclers for efficient DNA amplification. process, resulting in wasting the mixture and lowering the quality of
The performance of their device was compared to that of conventional amplification.
thermocyclers in terms of amplification efficiency, power consumption, In the proposed microfluidic chip, the target DNA sample and the
and duration. The results showed that PCB-based miniaturized ther­ master mix solution are first mixed inside the micromixer and streamed
mocycling achieves faster DNA amplification with significantly smaller directly into the reaction well. The reaction chamber well is located
power consumption while exhibiting similar efficiency to conventional above the microheater which carries out the thermocycling process
thermocyclers. Simulations are used to guide the design of such devices while the controller unit monitors the temperature inside the well dur­
and propose methods for further improvement of their performance. ing the PCR reaction. After PCR reaction is completed, the PCR product
Our proposed device employs gold electrodes etched on glass as a is analyzed by gel electrophoresis and the disposable part is replaced

2
H. Mashouf et al. Chemical Engineering and Processing - Process Intensification 189 (2023) 109394

Fig. 1. The microfluidic device layout for PCR amplification (a) Schematic of the integrated device, and (b) Exploded view of the microchip presenting inte­
gral components.

with a new one. 2.1.2. Temperature control unit


The cycling protocol in PCR reaction includes three stages: main­
2.1.1. Thermocycler design taining a steady temperature, cooling, and heating which are automat­
Temperature cycling is performed by heating and cooling the PCR ically mediated with the use of a microcontroller. Schematic of the
mixture according to a specific protocol. The thin film microheater temperature control unit is illustrated in Fig. S2 of supplementary
carries out the thermocycling process by Joule heating method. The document. The microheater is connected to the power supply using
temperature gradient is formed inside the PCR reaction well as electric crocodile connections (B&K Precision Corporation, Model 1672)
current passes through the thin film resistive electrode located beneath through the microcontroller. The power supply operates on constant
it. Achieving uniform temperature distribution over a designated area is voltage mode while the controller monitors the time of thermocycling
the prime concern in developing microheaters. The pattern design of the stages and temperature inside the reaction well. The programmable
electrode plays a crucial role in addressing this concern, and studies controller provides tuned voltage to the heater based on the feedback
have been made to find the efficient heater layout. Geometries with from the sensor as the temperature set points are altered according to the
meandering narrow line electrodes are generally incorporated since they algorithm.
reduce the power consumption [37,49]. Integrated thin film sensors were not an appropriate option for the
Several designs were tested in this study aiming to provide temper­ proposed device. Among temperature sensors, thermistors can offer the
ature uniformity with ±1 ◦ C precision across a circular area that houses required precision and stability while being affordable, compact, and
the PCR well. An infrared imager was used to investigate temperature compatible with microcontrollers. A 10KΩ thermistor (B3950 NTC,
uniformity. The designs also focused on providing fast thermal response Adafruit) with a negative temperature coefficient is used as the tem­
times while lowering the power required to heat the electrodes. In order perature sensor in the micro device. It is crucial to verify that the
to accomplish these goals, the size of the electrodes and the distance measured temperature correlates with the actual temperature distribu­
between them were gradually modified by conducting several experi­ tion of the PCR mixture in order to ensure an accurate and repeatable
ments until satisfactory results were achieved. amplification. The sensor was thermally calibrated to report the tem­
Fig. 2 shows the microheaters fabricated via photolithography by perature inside the well. The calibration of the temperature sensing unit
patterning Au on glass substrate. The edges of thin-film heaters always was performed by comparing the temperatures read by two thermistors,
show lower temperatures compared to the middle region because of low one embedded inside the reaction well, and another attached to the glass
ambient temperature. One approach that has been implemented to slide. In order to achieve the minimum temperature difference between
address this issue is to increase the resistance in the corners of the heater the two sensors (less than 0.1 ◦ C), two readings were compared for
[41]. However, this idea complicates the design of the pattern and its temperatures ranging from 40 ◦ C to 98 ◦ C with intervals of 2 ◦ C, and the
fabrication as well. Besides working on the pattern of the electrodes and sensor attached to the glass slide was adjusted accordingly. The method
their dimensions, microheaters cover a larger area than the reaction well of calibration is discussed in supplementary document.
to compensate for heat loss at the edges. In all proposed designs, the
heating region can be fitted into a square with a side length of 1 cm. The 2.1.3. Micromixer design
detailed dimensions can be found in Fig. S1. Mixing in the microscale often occurs as a result of molecular
diffusion process. The low molecule diffusivities would make fluid
mixing extremely challenging and the process of mixing through

Fig. 2. Fabricated microheaters by patterning gold on glass substrate. (a) Design 1, (b) Design 2, and (c) Design 3.

3
H. Mashouf et al. Chemical Engineering and Processing - Process Intensification 189 (2023) 109394

diffusion would take an excessive amount of time and require an transferring the heater pattern, the mask layout was designed using a
extremely lengthy channel [50,51]. Therefore, increasing the contact CAD program, and UV exposure was carried out for 8 s under a wave­
area between the fluids to be mixed is necessary to hasten the process length of 365 nm and 20 mW/cm2 (Model 204IR, OAI). Then, the
and reduce the mixing length. Several micromixers were fabricated and photoresist was developed for 45 s in developer (MICROPOSIT® MF®-
analyzed to find a convenient geometry for our application. Folding, 319). To perform the wet etching process, the glass slide with patterned
stretching, and splitting of the flow streams facilitate mixing at low photoresist was immersed for 30 s into a gold etchant (TFA gold etchant,
Reynolds numbers which can be generated by disturbing the flow with Transene). To remove the unwanted chrome layer, the glass was
creating curved channels or embedding obstacles throughout the immersed in chrome etchant (TFE chromium etchant, Transene) for 10 s.
channel. Fig. 3 presents the structure of two micromixer with their The residual photoresist was then stripped by MICROPOSIT™
detailed dimensions. These designs were selected and underwent mod­ REMOVER 1165. At the final stage, the glass was rinsed with water and
ifications to improve mixing quality by inserting sharp edges and isopropanol (IPA) and dried with nitrogen gas.
grooves in channel walls. The channel length is around 80 mm and 60 The steps for fabricating reaction wells are as follows: 10:1 wt ratio of
mm for design 1 and design 2, respectively. PDMS pre-polymer and curing agent (SYLGARD™ 184, Dow Inc., MI,
In this study, the mixing index is calculated by comparing the stan­ USA) was mixed. The mixture was poured into a mold and degassed for
dard deviation of the pixel intensity of the desired point to that in the 5 min, followed by baking for 2 h at 80 ◦ C in the oven to create PDMS
entirely unmixed section (inlet) [52]. To determine the standard devi­ sheets. These sheets with a thickness of 3 mm were then cut down into
ation, two fluid streams were flowed through the mixer. DI water was small blocks of 2 × 1.5 cm. Biopsy punches were used to make holes
injected from one inlet and a colorful stream from another one. The throughout the block to form the reaction well. The temperature uni­
experimental apparatus for evaluating mixing quality is shown in formity analysis is conducted to choose the appropriate well diameter.
Fig. 4a. ECHO Revolution microscope took RGB Images from the desired Negative photoresist SU-8 2075 (MicroChem, MA, USA) was used to
region inside the micromixer channel. These images were then con­ fabricate the master for replica molding of the micromixer. A 3-inch
verted to greyscale values via homemade code in computer software. silicon wafer was first cleaned using isopropanol (IPA). The photore­
These values range from 0 to 255, corresponding to the intensity of each sist was spin-coated on the wafer at a speed of 500 rpm for 10 s followed
pixel in the original image. The values were then normalized according by a speed of 2100 rpm for 30 s to create approximately 100 µm chan­
to the maximum intensity, resulting in normalized values Xi which nels. The wafer was then put onto a hot plate for 5 minutes at 65 ◦ C and
range from 0 to 1. The mixing index is then expressed by: 15 min at 95 ◦ C to perform the soft baking process. The mixer pattern
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
∑N was exposed onto the wafer by UV light (365 nm and 20 mW/cm2) with
a mask aligner for 13 s. Post-exposure baking processes were performed
1 2
σ i=1 (Xi − 〈X〉)
(1)
N
MI = 1 − = 1 − √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
∑N ̅
at 65 ◦ C and 95 ◦ C for 4 and 10 min, respectively. Then, the wafer was
σ0 1 2
i=1 (X0i − 〈X0 〉)
N developed in SU-8 developer for around 8 min, washed with isopropanol
Where, 〈X〉 denotes the average value over the normalized values, (IPA), and dried with nitrogen gas. For the last step of fabricating the
and N is the number of data. Standard deviation at the non-mixing mold master, the wafer underwent hard baking process for 5 min at
section (σ 0 ) would be equal to 0.5 if the flow rates are similar at the 150 ◦ C. After molds are prepared, micromixer PDMS blocks with 2 mm
inlet, implying completely segregated streams. This specific formula was thickness are created by implementing similar procedure as reaction
picked over the others because studies have demonstrated that it is least well blocks were fabricated.
impacted by lighting [53]. Fig. 4b demonstrates the process of calcu­ To fabricate the disposable part, fluidic access points were punched.
lating the mixing index. Then, reaction well block and micromixer block were exposed to oxygen
plasma for 20 s (Plasma Etch, PE-50, USA), brought into contact with
each other, and incubated at 80 ◦ C for 2 h to form an irreversible bond.
2.2. Fabrication The final stage of assembling the disposable part is bonding the block of
the reaction well and micromixer to a coverslip (22 × 22 mm, Fish­
The components of the microfluidic device are fabricated separately erbrand™) using oxygen plasma.
and then assembled together. To fabricate the microheater, glass slides
were first deposited with gold in the sputtering machine. The thickness 2.3. Sample preparation
of gold is 1000 Å, and 100 Å chrome acts as the adhesion layer. Positive
photoresist (MICROPOSIT™ S1813™ G2) was spin-coated on the slide The PCR reagents were purchased from Integrated DNA technolo­
at a spin speed of 500 rpm for 10 s and a spread speed of 3000 rpm for gies. The PCR master mix consists of 2 µL Deoxynucleoside triphosphate
20 s. The glass slide was then soft-baked at 115 ◦ C for 2 min. For (dNTP), 2μL 10X PCR buffer, 1 μL MgCl2, 10.8 μL nuclease-free water,

Fig. 3. Micromixer structures. (a) Design 1, and (b) Design 2.

4
H. Mashouf et al. Chemical Engineering and Processing - Process Intensification 189 (2023) 109394

Fig. 4. The experimental procedure for evaluating mixing efficiency (a) Experimental setup and fabricated micromixer structures, and (b) Mixing index calcula­
tion steps.

and 0.4 μL Taq gold (AmpliTaq Gold™ DNA Polymerase). Λ-phage DNA 3. Components performance
(Thermo Scientific™, USA) with concentration of 475.2 ng/μL was used
as template. The forward and reverse primers used were 5′ - ATG CCA 3.1. Microheater
CGT AAG CGA AAC A − 3′ , and 5′ - GCA TAA ACG AAG CAG TCG AGT
− 3′ , respectively, each with 0.4 μL volume. The master mix (17 µL) was 3.1.1. Spatial thermal uniformity
automatically mixed with of 100-fold diluted λ-phage DNA (3 µL) The microheaters are designed to create uniform heating and fast
through the micromixer and directly dispensed into the reaction well of thermal response with low power consumption. Temperature uniformity
the disposable block. To prevent evaporation, the sample was covered in an area can be defined as where the temperature difference between
with 10 μL of mineral oil. set point temperatures is less than 1 and/ or 2 ◦ C [54]. The spatial
The PCR thermal cycling was applied under the following protocol: thermal uniformity of the proposed designs was verified by an infrared
First, to activate the Taq polymerase predenaturation was done at 95 ◦ C imaging system, and their performance was compared over circular
for 5 min. Then, PCR cycle was repeated 40 times consisted of: dena­ areas. As gold directly reflects infrared light from its surroundings, an
turation of DNA template at 95 ◦ C for 15 s, annealing of primer at 54 ◦ C image taken with the resistor facing the IR camera would show a heat
for 20 s, and extension of the DNA at 72 ◦ C for 20 s. Final step was map of the surroundings rather than the actual temperature of the
elongation at 72 ◦ C for 1 min. Upon PCR completion, the amplified electrodes [37]. Therefore, images were taken from the backside of the
product was analyzed by gel electrophoresis in a 2% agarose gel Invi­ glass patterned with heater electrodes. Fig. 5 illustrates the images taken
trogen, and evaluated compared to the positive control band obtained by IR camera (FLIR Flip 4) from microheaters at the temperature of
by a conventional thermocycler (T100 Thermal Cycler, Bio-Rad, USA). 95 ◦ C (denaturing temperature) where Area1 is a circle of 4 mm in
Rnase-free water was used for the negative control as well. The resulting diameter.
DNA fragments size was measured by GeneRuler 100 bp DNA Ladder The pattern of heating maps in design 1 and design 3 are almost fully
(SM0241, Thermo Scientific™, USA). symmetric unlike rectangular-shaped heating map of design 2.
Numerous small nodes located at small distances from each other pre­
vents large heat dissipations in design 2, resulting in lowest temperature
deviation in Area 1 from the set point and not allowing the temperature
to drop significantly compared to other two designs. Design 1 reported
the most deviation due to the high thermal dissipation from its edges,
whereas in design 3 the center of the heater is surrounded by electrodes

Fig. 5. Spatial thermal uniformity comparison of microheater designs over a circular region using IR imaging at 95 ◦ C. (a) design 1, (b) design 2, and (c) design 3.

5
H. Mashouf et al. Chemical Engineering and Processing - Process Intensification 189 (2023) 109394

to avoid heat dissipation. With the help of these electrodes, design 3 The temperature uniformity was analyzed inside three circular areas
showed an average temperature of 94.9 ◦ C over measured points inside with diameters of 3.5, 4.5, and 5.5 mm at 95 ◦ C since highest deviation
Area 1. Evidently, as the circular area being studied gets bigger, the from desired setpoint happened at this temperature. These diameters are
average and minimum values drop. illustrated in Fig. 6a with D1, D2, and D3 annotations, respectively. We
Changing the voltage from 1 V to 1.8 V increased the temperature in considered the span of ±1 ◦ C temperature difference as uniform. The
design 2 from 54 ◦ C to 95 ◦ C. Although this might sound pretty satisfying uniformity is then reported as the percentage of measured points in the
for a heater, it might not work well for a thermocycler. A small alteration circular region which are inside this temperature span. The temperature
in voltage drastically changes the spatial temperature, making temper­ uniformity results for each diameter are illustrated in Fig. 6b. D3 shows
ature control challenging. Moreover, fast heating ramps do not allow unfavorable results with temperature uniformity of 68% at 95 ◦ C. D1, on
sufficient thermal distribution inside PCR mixture, impacting the quality the other hand, suggests superior performance by 100% uniformity at all
of amplification. In contrast, design 1 requires around 14 V to reach temperatures. D2 suggests uniformity of 93.1% which is pretty high as
95 ◦ C implying notable power consumption although showing adequate well, and it is not affected by temperature alterations. Nonetheless, the
thermal uniformity. Design 3 needed around 8.1 V and less than 2 Watts height of PCR mixture corresponding to D1 would be equal to 2.08 mm,
power for maintaining the temperature at 95 ◦ C making it suitable to be 66% higher than the height for D2. Furthermore, the total height of the
integrated into a microfluidic device. Same performance analysis on reaction well block will be affected as well since it must occupy mineral
various circular regions and different temperatures almost presented oil which will result in higher consumption of material (PDMS). How­
similar results on spatial thermal uniformity. Considering these expla­ ever, we decided to choose the final well diameter of 4 mm to have
nations, design 3 appeared to be the best fit for the microheater pattern. improved temperature uniformity. Also, the height difference between
Design 3 suggested similar heating map patterns with average tem­ D2 and a diameter of 4 mm is not that significant compared to D1.
peratures of 72.1 ◦ C and 54.1 ◦ C at extending and annealing tempera­ Table S1 reports the mixture heights corresponding to each reaction well
tures, respectively which is highly satisfying. The deviation of maximum diameter.
and minimum temperatures from the desired set point intensifies as the The plot presented in Fig. 6c shows the temperature along a hori­
temperature rises. This happens since heat dissipation to the surround­ zontal line crossing through the heated region to better understand
ings becomes easier and larger as the temperature difference with the temperature fluctuations inside the circle of D2 at a temperature of
ambient increases. The IR images of design 3 at 72 ◦ C and 54 ◦ C, can be 95 ◦ C. The maximum temperature is 95.8 ◦ C, as illustrated before in
found in Fig. S4 of the supplementary document. Fig. 5c. Large drops in the plot are probably due to the regions which are
not covered by the heater electrodes. The temperature decreases by
3.1.2. Reaction well diameter moving from the center of the heater toward the edges. The rate of drop
Fast thermal distribution and temperature uniformity inside the re­ in temperature apparently increases as we get closer to the edges of the
action well are key parameters in successful PCR process. As the diam­ heater.
eter gets smaller, temperature uniformity increases across the diameter,
but uniform thermal distribution along the well’s height might not 3.2. Dynamic temperature control unit
occur. Measuring temperature inside the reaction well at different
heights would be inaccurate due to the load effect of the external ther­ For successful PCR amplification, the temperature of the PCR
mometer. Therefore, it is tried to achieve a satisfactory compromise mixture must be carefully controlled and regulated. PCR reagents are
between height and temperature uniformity across the circular area of highly sensitive to temperature and the quality of amplification will be
the corresponding diameter. Fig. S5 illustrates the reaction chamber reflected if the control unit fails to maintain the temperature of the
PDMS block, its height, and reaction well diameter punched through it. mixture at specific set points according to the protocol with the lowest
Well diameters and corresponding mixture heights can be found in possible fluctuations. The control system in this work is programmed to
Table 1 of supplementary document as well. operate in the temperature range of common PCR protocols. Sufficient

Fig. 6. Well diameter analysis according to the selected microheater (design 3), (a) IR image at 95 ◦ C annotated with different well diameters, (b) Uniform area
percentage inside different well diameters with ±1 ◦ C accuracy, and (c) Cross-section temperature distribution along a horizontal line at 95 ◦ C.

6
H. Mashouf et al. Chemical Engineering and Processing - Process Intensification 189 (2023) 109394

number of tests were conducted to analyze the performance of the 3.3. Micromixer
control system.
54 ◦ C and 72 ◦ C are common PCR temperatures for annealing and At low Reynolds numbers, molecular diffusion prevails in the mixing
extending. Fig. 7a and 7b show the heating performance starting from mechanism of two streams. Therefore, the time the working fluid stays
these temperatures to two arbitrary setpoints followed by maintaining in the micromixer influences the mixing quality. The volume ratio of the
the temperature for around 15 s. According to the plots, the temperature target DNA to master mix is 3:17. Different methods can be implemented
fluctuations after reaching the desired set point is lower than 1 ◦ C, which to enhance the mixing efficiency of micromixers such as increasing the
can be regarded as uniform. Although microheater design and well length and height of the microchannels, changing the type and design of
diameter were selected in a way to facilitate thermal uniformity inside the micromixer, or increasing the hydrophilicity of PDMS by Oxygen
the well, the control system must also allow sufficient time for the re­ plasma treatment or surface coating. Efforts were made to enhance the
action well to reach uniform thermal distribution. To ensure the tem­ efficiency of the process by implementing methods that require fewer
perature uniformity of the sample inside the reaction well, a equipment and cost-efficient fabrication steps. Experiments were con­
temperature transition time is required for the set temperature to be ducted on micromixer structures to find out the appropriate design,
reached inside the well to ensure that the sample reaches the exact channel height, and flow rate for sufficient mixing efficiency. It is more
temperature as the sensor is recording. To ensure this smooth transition practical to report the amplification results in mixing efficiency ranges
of the temperature, the controlling system dynamically adjusts the rather than a minimum value for mixing index required for successful
heating and cooling rate by considering the aforementioned criteria. amplification. We found that for efficiencies below 0.85, no amplifica­
Considering all the limitations, the applied heating rates ranged be­ tion occurs, or only unspecific bands appear. For efficiencies between
tween 2 and 2.6 ◦ C/s. As a result of a more uniform temperature dis­ 0.85 and 0.9, the results were not repetitive, and the brightness of bands
tribution inside the well at lower temperatures, higher heating rates are were not always similar to the positive control, while in efficiencies
observed in the well at lower temperatures. In addition, as the difference above 0.9, results were satisfying. E-gel electrophoresis test results can
between the two set points narrows, the heating rate decreases. be found in Fig. S6 of supplementary document.
Fig. 8a and b display the cooling performance of the device in two Fig. 9a presents the effect of channel height on mixing efficiency at
separate plots. The set points for each plot are different. Here, the per­ total flow rate of 20 µL.min− 1. Experiments conducted at other total flow
formance of the device is compared to when fans are employed as well. rates between 0.5 and 50 µL.min− 1 followed the same trend. Increasing
Two fans (Delta Electronics, AUC0512, 20.2 CFM) were placed at the channel height will improve the mixing quality since the contact area
bottom of the microheater to facilitate cooling. The average cooling between the two streams and the fluid-fluid interaction time increases;
rates are smaller than the heating rates. It is apparent that fan installa­ hence the diffusion rate intensifies as well. Channel height of 40 µm
tion results in an increase in the cooling rate. The maximum cooling suggested poor mixing quality. Significant improvement in mixing
rates were observed in the absence of a fan at 2.1 ◦ C/s, and with a fan at quality was observed by increasing channel height from 40 µm to 100
1.7 ◦ C/s. The protocol chosen for this particular application involved µm although the improvement is not notable from 100 µm to 180 µm.
denaturation and annealing stages. Considering the cooling rate differ­ From the practical view, channels with high heights are difficult to
ence between using and not using fans we decided against including it in fabricate, and the molds are more susceptible to getting deformed due to
our device. Utilizing fans decreases cycle time by around 5 s which is not their high aspect ratio. Therefore, channel height of 100 µm is consid­
that significant. However, it would be an extra component that not only ered for the micromixers.
increases the size of the whole microfluidic device but also increases its The adverse effect of increasing the flow rate at a constant length is
power consumption. expected since the working fluids will have less time to diffuse and are
driven toward the outlet without having sufficient time to interact. The
bar chart in Fig. 9b presents the effect of total flow rate on mixing ef­
ficiency. Both designs suggested high efficiencies while total flow rate of

Fig. 7. Heating and temperature maintaining performance. (a) Heating from 54 ◦ C to 72 ◦ C and 90 ◦ C, (b) Heating from 72 ◦ C to 85 ◦ C and 95 ◦ C.

7
H. Mashouf et al. Chemical Engineering and Processing - Process Intensification 189 (2023) 109394

Fig. 8. Comparison of cooling and temperature maintaining performance of the microheater by utilizing fans (a) Cooling from 95 ◦ C to 54 ◦ C, and (b) Cooling from
70 ◦ C to 50 ◦ C.

Fig. 9. Micromixers mixing performance comparison. (a) Mixing efficiency at different channel heights and total flow rate of 20 µL.min− 1 (b) Mixing efficiency at
different total flow rates and channel height of 100 µm, and (c) Mixing efficiency at channel height of 100 µm and total flow rate of 10 µL.min− 1 with respect to
channel length.

20 µL.min− 1 showed poor performance in comparison to other flow min− 1 and lower takes too much time. Since the mixing efficiency of
rates. In addition to poor performance, a high flow rate might damage total flow rate at 10 µL.min− 1 is sufficient for our application, we
the bonding between the micromixer PDMS channel and reaction decided to choose this flow rate over flow rate of 4 µL.min− 1. Overall,
chamber PDMS block, resulting in the leakage of the PCR mixture and design 1 displayed better performance over design 2 in terms of final
impeding the PCR reaction. Handling low target DNA volume (3 µL) can mixing indices. This can be explained with the chaotic flow regime
also be difficult. Therefore, it was decided not to choose a total flow rate created by the high number of grooves across its channel walls.
of 20 µL.min− 1 and higher. Both total flow rates of 10 and 4 µL.min− 1 Although design 1 achieved better mixing efficiency, design 2 is compact
suggested satisfactory results. However, mixing at total flow rate of 4 µL. and more convenient for a small disposable device. Unlike design 2,

8
H. Mashouf et al. Chemical Engineering and Processing - Process Intensification 189 (2023) 109394

design 1 occupies a larger space making its production costly as it re­ amplification process takes around 70 min, which is lower than con­
quires more material. ventional PCRs.
It should be noted that the mixing channel of design 1 is around 20 The amplification performance of the microfluidic PCR device was
mm longer than design 2. According to the plot in Fig. 9c, they show a verified by a conventional PCR referred to as λ DNA positive control.
very close mixing index at a distance of 60 mm from the inlet, which is Fig. 12b presents the final results visualized by E-gel electrophoresis.
the total channel length of design 2. Design 1 reaches high efficiencies Lane M is the 100-bp ladder DNA marker. Signal in Lane 1 is obtained by
much faster than design 2 due to its special obstacle-based geometry. using the conventional PCR machine for positive control, and lane 5
However, it was not possible to make design 1 as compact as design 2, represents negative control. Bands in lanes 2, 3, and 4 are the PCR
even by reducing its length. The most challenging issue, however, was products obtained by different microfluidic devices. To ensure that
the formation of bubbles in several regions around the grooves, which contamination is prevented, and PCR amplification results are repro­
can impact the mixing quality or even block the channel, which did not ducible, experimental tests were conducted on microfluidic devices with
happen in experiments using design 2. different disposable and reusable parts. The results demonstrate that this
Since total flow rate is decided to be 10 µL.min− 1, the flow rates at system can generate the correct amplified PCR base pair and that
the inlets should match with volume ratio of PCR mixtures. Therefore, degradation does not occur throughout the process. The brightness of
the flow rate at one of the inlets would be 1.5 µL.min− 1 and 8.5 µL.min− 1 bands can be used to qualitatively assess the efficiency of amplification.
in the other one. Fig. 10a demonstrate microscope images at different Since no significant difference in the PCR band intensities was observed,
regions throughout the selected micromixer (design 2) under this flow it can be interpreted that the proposed device can perform the amplifi­
rate and channel height of 100 µm. The continuous growth in mixing cation with high efficiency.
quality along the channel is noticeable after the working fluids pass
through each loop. Plots illustrating mixing performance at three 5. Conclusion
different heights and three different total flow rates are illustrated in
Fig. 10b to better understand the mixing indices. The regions used to In this study, a disposable, compact, low-cost and easy-to-fabricate
calculate mixing indices have identical distances from each other and microfluidic device is designed to amplify λ-DNA. The disposable
correspond to the microscope images in Fig. 10a. The growth rate of feature of the device eliminates the run-to-run carryover contamination,
mixing quality decreases as the mixing efficiency rises to high values. while its integration with reusable heating system reduces fabrication
Mixing efficiency of 0.8 is surpassed after the flow reaches halfway costs. The need for integrating a microvalve to isolate the PCR mixture
through the mixer at selected characteristics, and the final efficiency is from flowing back to micromixer is eliminated by the layered design of
equal to 94.17%. the disposable part. The microheater is fabricated by patterning gold
electrodes on glass substrate and carries out the thermocycling process.
4. Integrated device performance Around 95% of the area housing the reaction well have temperature
uniformity with thermal variation of <1 ◦ C at protocol setpoints (54 ◦ C,
The developed compact microfluidic chip is shown in Fig. 11a. The 72 ◦ C, 95 ◦ C). The microcontroller unit can maintain the temperature
heating component of the chip can be reused for future amplifications, with less than 1 ◦ C error inside common working range of PCR protocols.
while the disposable section is simply replaced with a new one after each The developed passive micromixer suggested mixing efficiency of
PCR cycle is finished. The experimental setup during PCR reaction is 94.17% at channel height of 100 µm and total flow rate of 10 µL.min− 1.
shown in Fig. 11b. After the protocol is finished, the amplified mixture is The integrated device proposed high-quality amplification results
recovered for the post PCR process using E-gel electrophoresis. comparable to the conventional PCR. Additionally, the studies on the
The PCR reaction is performed following a protocol consisting of 40 reproducibility of the device shows its superior performance in gener­
cycles. The performance of the developed microfluidic device is pre­ ating consistent and reliable results for different amplification protocols.
sented in Fig. 12a. The temperature was meticulously maintained at PCR Further study can be focused on employing multiple microheaters on a
step set points with a deviation of less than 1 ◦ C/s. The whole single chip in order to increase the throughput and utilization of the

1
Fig. 10. Micromixer performance. (a) Microscope images throughout design 2 at total flow rate of 10 µL.min− and channel height of 100 µm, and (b) Performance of
design 2 at different total flow rates and channel heights.

9
H. Mashouf et al. Chemical Engineering and Processing - Process Intensification 189 (2023) 109394

Fig. 11. Integrated microfluidic device. (a) Fabricated microfluidic chip, and (b) Experimental setup for conducting PCR reaction.

Fig. 12. Integrated device performance. (a) 15 PCR cycles using the developed microfluidic device, and (b) E-gel results of the developed microfluidic device
compared to conventional PCR.

suggested device for the amplification of the target gene from an ani­ Investigation, Validation. Sepideh Pakpour: Conceptualization, Re­
mal’s total DNA. sources, Funding acquisition. Mina Hoorfar: Conceptualization, Su­
pervision, Funding acquisition, Resources.
Statement of novelty and significance
Declaration of Competing Interest
Rapid heat transfer, and low sample and energy consumption are
advantages of microfluidic PCR devices. However, they are unsuitable The authors declare that they have no known competing financial
for mass production and portable use. Our proposed compact device interests or personal relationships that could have appeared to influence
comprises several components integrated in a layered design aiming to the work reported in this paper.
reduce the fabrication cost and eliminate the chance of sample-to-
sample cross-contamination. The microheater and control system form Data availability
the reusable parts of the chip while the reaction well and micromixer are
disposable. The disposable element is easy to fabricate and replaced Data will be made available on request.
with a new one upon completion of PCR reaction. The performance of
the device is verified by amplifying lambda DNA samples.
Acknowledgment
CRediT authorship contribution statement
This work was supported by the Natural Science and Engineering
Hirad Mashouf: Investigation, Data curation, Writing – original Research Council (NSERC) of Canada under the Discovery Grant (Grant
draft, Writing – review & editing. Bahram Talebjedi: Conceptualiza­ no. 341873).
tion, Supervision, Investigation, Methodology, Software, Writing – re­
view & editing. Nishat Tasnim: Writing – review & editing, Project
administration. Maia Tan: Investigation, Validation. Sahar Alousi:

10
H. Mashouf et al. Chemical Engineering and Processing - Process Intensification 189 (2023) 109394

Supplementary materials [24] F. Cui, et al., Design and experiment of a PDMS-based PCR chip with reusable
heater of optimized electrode, Microsyst. Technol. 23 (8) (2017) 3069–3079,
https://doi.org/10.1007/s00542-016-3064-3.
Supplementary material associated with this article can be found, in [25] B. Talebjedi, M. Ghazi, N. Tasnim, S. Janfaza, M. Hoorfar, Performance
the online version, at doi:10.1016/j.cep.2023.109394. optimization of a novel passive T-shaped micromixer with deformable baffles,
Chem. Eng. Process. - Process Intensification 163 (2021), 108369, https://doi.org/
10.1016/J.CEP.2021.108369.
References [26] J. Jie, et al., Portable and battery-powered PCR device for DNA amplification and
fluorescence detection, Sensors 20 (9) (2020) 2627, https://doi.org/10.3390/
[1] C. Zhang, D. Xing, Miniaturized PCR chips for nucleic acid amplification and s20092627.
analysis: latest advances and future trends, Nucleic Acids Res. 35 (13) (2007) [27] K.E. Herold, N. Sergeev, A. Matviyenko, A. Rasooly, Rapid DNA amplification using
4223–4237, https://doi.org/10.1093/NAR/GKM389. a battery-powered thin-film resistive thermocycler, Methods Mol. Biol. 504 (2009)
[2] T. Vilkner, D. Janasek, A. Manz, Micro total analysis systems. Recent 441–458, https://doi.org/10.1007/978-1-60327-569-9_24/FIGURES/6_24.
developments, Anal. Chem. 76 (12) (2004) 3373–3386, https://doi.org/10.1021/ [28] R.A. Mendoza-Gallegos, A. Rios, J.L. Garcia-Cordero, An affordable and portable
AC040063Q. thermocycler for real-time PCR made of 3D-printed parts and off-the-shelf
[3] P.A. Auroux, Y. Koc, A. DeMello, A. Manz, P.J.R. Day, Miniaturised nucleic acid electronics, Anal. Chem. 90 (9) (2018) 5563–5568, https://doi.org/10.1021/ACS.
analysis, Lab Chip 4 (6) (2004) 534–546, https://doi.org/10.1039/B408850F. ANALCHEM.7B04843/SUPPL_FILE/AC7B04843_SI_001.PDF.
[4] H. Zhu, H. Zhang, S. Ni, M. Korabečná, L. Yobas, P. Neuzil, The vision of point-of- [29] S. Jeong, et al., Portable low-power thermal cycler with dual thin-film Pt heaters
care PCR tests for the COVID-19 pandemic and beyond, TrAC - Trends Anal. Chem. for a polymeric PCR chip, Biomed. Microdevices 20 (1) (2018) 14, https://doi.org/
130 (2020), https://doi.org/10.1016/j.trac.2020.115984. 10.1007/s10544-018-0257-9.
[5] L. Syedmoradi, M. Daneshpour, M. Alvandipour, F.A. G.omez, H. Hajghassem, [30] D. Moschou and A. Tserepi, “Lab on a Chip CRITICAL REVIEW The lab-on-PCB
K. Omidfar, Point of care testing: the impact of nanotechnology, Biosens. approach: tackling the μTAS commercial upscaling bottleneck,” vol. 17, p. 1388,
Bioelectron. 87 (2017) 373–387, https://doi.org/10.1016/J.BIOS.2016.08.084. 2017, doi: 10.1039/c7lc00121e.
[6] Y. Zhang, P. Ozdemir, Microfluidic DNA amplification–a review, Anal. Chim. Acta [31] H. Nagai, Y. Murakami, K. Yokoyama, E. Tamiya, High-throughput PCR in silicon
638 (2) (2009) 115–125, https://doi.org/10.1016/j.aca.2009.02.038. based microchamber array, Biosens. Bioelectron. 16 (9–12) (2001) 1015–1019,
[7] M.R. G.reen, J. Sambrook, Polymerase chain reaction, Cold Spring Harb. Protoc. https://doi.org/10.1016/S0956-5663(01)00248-2.
2019 (6) (2019), https://doi.org/10.1101/PDB.TOP095109. [32] K. Hu, Z. Chen, J. Huang, Research on temperature measuring positions selection
[8] C.Y. Lee, L.M. Fu, Recent advances and applications of micromixers, Sens. and verification for polymerase chain reaction instruments, in: Proceedings - 2011
Actuators B Chem. 259 (2018) 677–702, https://doi.org/10.1016/j. 4th International Conference on Biomedical Engineering and Informatics, BMEI
snb.2017.12.034. 2011 3, 2011, pp. 1165–1169, https://doi.org/10.1109/BMEI.2011.6098553.
[9] E. Samiei, M. Tabrizian, M. Hoorfar, A review of digital microfluidics as portable [33] M.G. Roper, C.J. Easley, L.A. Legendre, J.A.C. Humphrey, J.P. Landers, Infrared
platforms for lab-on a-chip applications, Lab Chip 16 (13) (2016) 2376–2396, temperature control system for a completely noncontact polymerase chain reaction
https://doi.org/10.1039/C6LC00387G. in microfluidic chips, Anal. Chem. 79 (4) (2007) 1294–1300, https://doi.org/
[10] B. Talebjedi, M. Heydari, E. Taatizadeh, N. Tasnim, I.T.S. Li, M. Hoorfar, Neural 10.1021/AC0613277/SUPPL_FILE/AC0613277SI20060809_114422.PDF.
network-based optimization of an acousto microfluidic system for submicron [34] J. El-Ali, I.R. Perch-Nielsen, C.R. Poulsen, D.D. Bang, P. Telleman, A. Wolff,
bioparticle separation, Front. Bioeng. Biotechnol. 10 (2022) 1–15, https://doi.org/ Simulation and experimental validation of a SU-8 based PCR thermocycler chip
10.3389/fbioe.2022.878398. with integrated heaters and temperature sensor, Sens. Actuators A Phys. 110 (1–3)
[11] R. Abedini-Nassab, J. Wirfel, B. Talebjedi, N. Tasnim, M. Hoorfar, Quantifying the (2004) 3–10, https://doi.org/10.1016/j.sna.2003.09.022.
dielectrophoretic force on colloidal particles in microfluidic devices, Microfluid. [35] T.-M. Hsieh, C.-H. Luo, F.-C. Huang, J.-H. Wang, L.-J. Chien, G.-B. Lee,
Nanofluid. 26 (5) (2022) 1–11, https://doi.org/10.1007/S10404-022-02544-0/ Enhancement of thermal uniformity for a microthermal cycler and its application
FIGURES/7. for polymerase chain reaction☆, Sens. Actuators B Chem. 130 (2) (2008) 848–856,
[12] B. Talebjedi, A. Abouei Mehrizi, B. Talebjedi, S.S. Mohseni, N. Tasnim, M. Hoorfar, https://doi.org/10.1016/j.snb.2007.10.063.
Machine learning-aided microdroplets breakup characteristic prediction in flow- [36] E.T. Lagally, C.A. Emrich, R.A. Mathies, Fully integrated PCR-capillary
focusing microdevices by incorporating variations of cross-flow tilt angles, electrophoresis microsystem for DNA analysis, Lab Chip 1 (2) (2001) 102, https://
Langmuir 38 (34) (2022) 10465–10477, https://doi.org/10.1021/acs. doi.org/10.1039/b109031n.
langmuir.2c01255. [37] H.-W. Veltkamp, F. Akegawa Monteiro, R. Sanders, R. Wiegerink, J. Lötters,
[13] B. Talebjedi, N. Tasnim, M. Hoorfar, G.F. Mastromonaco, M. De Almeida Monteiro Disposable DNA amplification chips with integrated low-cost heaters †,
Melo Ferraz, Exploiting microfluidics for extracellular vesicle isolation and Micromachines (Basel) 11 (3) (2020) 238, https://doi.org/10.3390/mi11030238.
characterization: potential use for standardized embryo quality assessment, Front. [38] C. Fang, F. Ji, Z. Shu, D. Gao, Determination of the temperature-dependent cell
Vet. Sci. 7 (2021) 1139, https://doi.org/10.3389/fvets.2020.620809. membrane permeabilities using microfluidics with integrated flow and
[14] “MIT open access articles ultrafast, low-power, PCB manufacturable, continuous- temperature control, Lab Chip 17 (5) (2017) 951–960, https://doi.org/10.1039/
flow microdevice for DNA amplification,” 2019, doi: 10.1007/s00216-019-01911 C6LC01523A.
-1. [39] J. Martinez-Quijada, et al., Deterministic design of thin-film heaters for precise
[15] D. Moschou, et al., All-plastic, low-power, disposable, continuous-flow PCR chip spatial temperature control in lab-on-chip systems, J. Microelectromech. Syst. 25
with integrated microheaters for rapid DNA amplification, Sens. Actuators B Chem. (3) (2016) 508–516, https://doi.org/10.1109/JMEMS.2016.2536561.
199 (2014) 470–478, https://doi.org/10.1016/J.SNB.2014.04.007. [40] Z.Y. Wu, K. Chen, B.Y. Qu, X.X. Tian, X.J. Wang, F. Fang, A thermostat chip of
[16] N.Y. Lee, A review on microscale polymerase chain reaction based methods in indium tin oxide glass substrate for static polymerase chain reaction and in situ real
molecular diagnosis, and future prospects for the fabrication of fully integrated time fluorescence monitoring, Anal. Chim. Acta 610 (1) (2008) 89–96, https://doi.
portable biomedical devices, Microchimica Acta 185 (6) (2018) 285, https://doi. org/10.1016/j.aca.2007.12.044.
org/10.1007/s00604-018-2791-9. [41] T.-M. Hsieh, C.-H. Luo, J.-H. Wang, J.-L. Lin, K.-Y. Lien, G.-B. Lee, A two-
[17] M.B. Kulkarni, S. Goel, Advances in continuous-flow based microfluidic PCR dimensional, self-compensated, microthermal cycler for one-step reverse
devices—a review, Eng. Res. Express 2 (4) (2020), 042001, https://doi.org/ transcription polymerase chain reaction applications, Microfluid. Nanofluid. 6 (6)
10.1088/2631-8695/abd287. (2009) 797–809, https://doi.org/10.1007/s10404-008-0353-x.
[18] W. Liu, et al., Establishment and evaluation of a 30-minute detection method for [42] J. Wu, R. Kodzius, K. Xiao, J. Qin, W. Wen, Fast detection of genetic information by
SARS-CoV-2 nucleic acid using a novel ultra-fast real-time PCR instrument, an optimized PCR in an interchangeable chip, Biomed. Microdevices 14 (1) (2012)
J. Thorac. Dis. 13 (12) (2021) 6866–6875, https://doi.org/10.21037/JTD-21- 179–186, https://doi.org/10.1007/s10544-011-9595-6.
1288. [43] F.-C. Huang, C.-S. Liao, G.-B. Lee, An integrated microfluidic chip for DNA/RNA
[19] J.J. Chen, Z.H. Lin, Fabrication of an oscillating thermocycler to analyze the canine amplification, electrophoresis separation and on-line optical detection,
distemper virus by utilizing reverse transcription polymerase chain reaction, Electrophoresis 27 (16) (2006) 3297–3305, https://doi.org/10.1002/
Micromachines (Basel) 13 (4) (2022) 600, https://doi.org/10.3390/MI13040600/ elps.200600458.
S1. [44] D. Resnik, D. Vrtačnik, M. Možek, B. Pečar, S. Amon, Experimental study of heat-
[20] O. Frey, S. Bonneick, A. Hierlemann, J. Lichtenberg, Autonomous microfluidic treated thin film Ti/Pt heater and temperature sensor properties on a Si
multi-channel chip for real-time PCR with integrated liquid handling, Biomed. microfluidic platform, J. Micromech. Microeng. 21 (2) (2011), 025025, https://
Microdevices 9 (5) (2007) 711–718, https://doi.org/10.1007/s10544-007-9080-4. doi.org/10.1088/0960-1317/21/2/025025.
[21] G.A. Nasser, A.L. Abdel-Mawgood, A.A. Abouelsoud, H. Mohamed, S. Umezu, A.M. [45] N. Beyor, L. Yi, S.S. Tae, R.A. Mathies, Integrated capture, concentration, PCR, and
R.F. El-Bab, New cost effective design of PCR heating cycler system using Peltier capillary electrophoretic analysis of pathogens on a chip, Anal. Chem. 81 (9)
plate without the conventional heating block, J. Mech. Sci. Technol. 35 (7) (2021) (2009) 3523, https://doi.org/10.1021/AC900060R.
3259–3268, https://doi.org/10.1007/S12206-021-0646-5. [46] G.D. Kaprou, V. Papadopoulos, C.M. Loukas, G. Kokkoris, A. Tserepi, Towards PCB-
[22] R. Khnouf, M.A.K. Jaradat, D. Karasneh, F. Al-Shami, L. Sawaqed, B.A. A.lbiss, based miniaturized thermocyclers for DNA amplification, Micromachines (Basel)
Simulation and optimization of a single heater convective PCR chip and its 11 (3) (2020), https://doi.org/10.3390/mi11030258.
controller for fast salmonella enteritidis detection, IEEE Sens. J. 20 (22) (2020) [47] D.R. Reyes, D. Iossifidis, P.A. Auroux, A. Manz, Micro total analysis systems. 1.
13186–13195, https://doi.org/10.1109/JSEN.2020.3004285. Introduction, theory, and technology, Anal. Chem. 74 (12) (2002) 2623–2636,
[23] D. Han, et al., MCU based real-time temperature control system for universal https://doi.org/10.1021/AC0202435.
microfluidic PCR chip, Microsyst. Technol. 20 (3) (2014) 471–476, https://doi. [48] C.-S. Liao, Miniature RT-PCR system for diagnosis of RNA-based viruses, Nucleic
org/10.1007/S00542-013-1970-1/FIGURES/4. Acids Res. 33 (18) (2005) e156, https://doi.org/10.1093/nar/gni157.

11
H. Mashouf et al. Chemical Engineering and Processing - Process Intensification 189 (2023) 109394

[49] P. Bhattacharyya, Technological journey towards reliable microheater [52] L. Wang, S. Ma, X. Wang, H. Bi, X. Han, Mixing enhancement of a passive
development for MEMS gas sensors: a review, IEEE Trans. Device Mater. Reliab. 14 microfluidic mixer containing triangle baffles, Asia-Pacific J. Chem. Eng. 9 (6)
(2) (2014) 589–599, https://doi.org/10.1109/TDMR.2014.2311801. (2014) 877–885, https://doi.org/10.1002/apj.1837.
[50] W. Raza, S. Hossain, K.Y.Y. Kim, A review of passive micromixers with a [53] A. Hashmi, J. Xu, On the quantification of mixing in microfluidics, J. Lab. Autom.
comparative analysis, Micromachines (Basel) 11 (5) (2020) 455, https://doi.org/ 19 (5) (2014) 488–491, https://doi.org/10.1177/2211068214540156/ASSET/
10.3390/mi11050455. IMAGES/LARGE/10.1177_2211068214540156-FIG2.JPEG.
[51] N.-T. Nguyen, Z. Wu, Micromixers—a review, J. Micromech. Microeng. 15 (2) [54] D.J. Sadler, R. Changrani, P. Roberts, C.F. Chou, F. Zenhausern, Thermal
(2005) R1–R16, https://doi.org/10.1088/0960-1317/15/2/R01. management of bioMEMS: temperature control for ceramic-based PCR and DNA
detection devices, IEEE Trans. Compon. Pack. Technol. 26 (2) (2003) 309–316,
https://doi.org/10.1109/TCAPT.2003.815093.

12

You might also like