You are on page 1of 8

Electrophoresis 2006, 27, 5051–5058 5051

Zhenyu Wang1 Research Article


Andrea Sekulovic1, 2
Jörg P. Kutter1
Dang D. Bang3 Towards a portable microchip system with
Anders Wolff1
integrated thermal control and polymer
1
MIC – Department of Micro and
Nanotechnology, waveguides for real-time PCR
Technical University of Denmark,
Lyngby, Denmark A novel real-time PCR microchip platform with integrated thermal system and polymer
2
Department of Biotechnology,
Technical University of Delft, waveguides has been developed. The integrated polymer optical system for real-time
Delft, The Netherlands monitoring of PCR was fabricated in the same SU-8 layer as the PCR chamber, without
3
Department of Poultry, additional masking steps. Two suitable DNA binding dyes, SYTOX Orange and
Fish and Fur Animals, TO-PRO-3, were selected and tested for the real-time PCR processes. As a model,
Danish Institute for Food and
Veterinary Research, cadF gene of Campylobacter jejuni has been amplified on the microchip. Using the
Aarhus, Denmark integrated optical system of the real-time PCR microchip, the measured cycle thresh-
old values of the real-time PCR performed with a dilution series of C. jejuni DNA tem-
plate (2 to 200 pg/mL) could be quantitatively detected and compared with a conven-
Received June 8, 2006
tional post-PCR analysis (DNA gel electrophoresis). The presented approach provided
Revised August 28, 2006
reliable real-time quantitative information of the PCR amplification of the targeted
Accepted August 31, 2006
gene. With the integrated optical system, the reaction dynamics at any location inside
the micro reaction chamber can easily be monitored.

Keywords: Integrated thermal system / Polymer waveguides / Real-time PCR / SU-8


DOI 10.1002/elps.200600355

1 Introduction the final amplified product concentration often has large


variations caused by minor disturbances, e.g. reaction
During the last decade, micro-total analysis system (mTAS) components, thermal cycling fluctuation, or primer mis-
devices have had a remarkable impact on biochemical, alignment [3]. A solution to this problem is to use real-time
chemical, and pharmaceutical research activities, because PCR. In real-time PCR, the product formation is meas-
of their many potential advantages, such as reduced cost, ured during the reaction by using different fluorescent
portability, and low reagent consumption [1]. To realize DNA binding dyes or different types of fluorescently
such mTAS devices, the integration of different detectors to labeled probes such as hybridization probes [4], hydro-
monitor various parameters within the system is a crucial lysis probes [5, 6], and hairpin probes [7–11], which bind
step. Recently, we have demonstrated the application of an specifically to DNA targets. The DNA binding dye most
integrated polymer optical system for cell detection in a often used in real-time PCR is SYBR green I [12, 13]. This
micro flow cytometer [2]. The integrated optical system dye has very low fluorescence in solution, but its fluores-
provided more integration feasibility, higher capacity, and cence increases 1000-fold when it binds to dsDNA and
more precise alignment for various microsystems in com- therefore acts as a nonspecific fluorescent reporter of the
parison to the bulk optical system. In this study, we apply dsDNA concentrations. Measuring and plotting the fluo-
integrated optics for real-time monitoring of PCR. rescence intensities after each PCR cycle can establish a
PCR amplification curve. The cycle threshold value (CT) is
PCR is an enzyme-catalyzed nucleotide amplification
the cycle number at which the fluorescence intensity is
technique, routinely used in many different fields for
higher than the detection baseline level. This parameter
genetic identification. Although PCR is a robust and pre-
provides more accurate and real-time quantitative infor-
dictable method, quantification can be difficult because
mation on the PCR amplification process [14].

Correspondence: Dr. Anders Wolff, MIC – Department of Micro and PCR has also been realized on microchips. To date, three
Nanotechnology, Technical University of Denmark, Bldg 345 east, different types of PCR microchips have been developed,
DK-2800, Kgs. Lyngby, Denmark
featuring either a chamber [15–29], a continuous flow [30–
E-mail: aw@mic.dtu.dk
Fax: 145-45887762
33], or a droplet oscillation design [34]. Until now, only few
real-time PCR microsystems have been reported [35–37].
Abbreviation: CT, cycle threshold value Recently, Gulliksen et al. [36, 37] described a novel real-

© 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com


5052 Z. Wang et al. Electrophoresis 2006, 27, 5051–5058

time PCR microchip. However, this real-time PCR micro- cal elements. The PCR microchip was tested for real time
chip required an external optical system to detect the flu- PCR detection of the cadF gene of C. jejuni using DNA-
orescent probes. binding dyes.

We have applied the PCR microchips for the detection of


Campylobacter. Campylobacter jejuni is the most com-
mon food-borne bacterial pathogen that causes gastro- 2 Materials and methods
enteritis in humans [38]. Several groups have developed
different conventional real-time PCR procedures for rapid 2.1 Model of the heater array
detection of Campylobacter [39–41]. Previously, we
reported on a PCR microchip with integrated thermal Based on a previously integrated thermal system design
system for fast thermocycling [42] and this PCR micro- [42], the Pt heater array was remodeled and redesigned.
chip had been applied to detect C. jejuni. As an important A 2-D heat-transport model has been established in
step towards a portable genetic analytical microsystem, FEMLAB 3.1 (Fig. 1A). The heat is generated by the inte-
in this paper we present a novel real-time PCR microchip grated Pt heater array. The chip is passively cooled by
with integrated heater, thermometer and polymeric opti- heat conduction through the substrate to an aluminum

Figure 1. (A) Simulated thermal profile using a 2-D heat-


transfer simulation model in FEMLAB 3.1. (B) Comparison of
the temperature profiles inside the reaction chamber for two
different heat array designs.

© 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com


Electrophoresis 2006, 27, 5051–5058 Miniaturization 5053

heat sink and by natural convection from the lid. There is


extra heat loss at the edges of the chamber. To avoid the
detrimental effect of a “cold wall”, the integrated Pt heater
array was designed to generate a 1.1 times higher heat
density at the edges by reducing either the heater width or
the distance between the heaters, and by extending the
heaters outside of the chamber. The simulation results
(Fig. 1B) show a much better temperature profile for the
non-uniform heat source than for the uniform one. The
homogeneous temperature area (within 60.57C variation
at 947C) inside the chamber has been expanded three
times (from 16 to 49 mm2) in comparison to the previous Figure 2. Layout sketch of the chip with integrated
design [42]. waveguides and thermal system. The thermal system
consisted of an array of 104 Pt heaters and a thermo-
meter integrated beneath a 5 mm thick SU-8 protection
2.2 Chip design, fabrication and packaging layer. The dimension of the reaction chamber is
8 mm68 mm60.4 mm, to form a 25-mL PCR reactor. On
The fabrication of the real-time PCR microchip the microchip, four waveguides are placed inside the
(18 mm618 mm) is a three-mask process as described chamber for fluorescence detection. All the integrated
optical elements (waveguides, couplers) are defined in
previously [42]. First, the electrodes (100-Å Ti, 200 nm Pt)
the same SU-8 layer as the reaction chamber in one
for the integrated heater array and thermometer were photolithography step.
deposited on a 500 mm Pyrex substrate (Schott Corpora-
tion, Germany) by e-beam evaporation and defined in a
standard lift-off process. Secondly, on top of the metal Devices) were used to connect the four wide pads for the
layer, a 5 mm SU-8 (XP2005, MicroChem, USA) protection heater array for preventing current overloading at the
layer was fabricated to serve as the chamber floor. Finally, connecting areas. A LabVIew Proportional-Integral-Dif-
the 8 mm68 mm reaction chamber and optical systems ferential (PID) temperature-controlling program was used
were defined by a standard photolithography in a 400 mm to control the power of the integrated thermal system
thick SU-8 (XP2075, MicroChem) layer (refractive index using a 15 W custom-built power supply.
n = 1.59). The PCR chamber was designed with a relative
large volume (25 mL) so that the PCR reaction product
could be analyzed off-chip for comparison using conven- 2.3 Primers and real-time PCR conditions
tional methods. The chip structure is shown in Fig. 2.
Bacterial chromosomal DNA was isolated from an over-
The reaction chamber was hermetically sealed by a night culture of C. jejuni on blood agar plates incubated at
1 mm thick poly(dimethylsiloxane) (PDMS) lid with a 427C under micro-aerobic conditions (6% O2, 6% CO2,
refractive index n = 1.4, which also provided top cladding 4% H2, and 84% N2) as previously described [42]. The
for the waveguides. The substrate (n = 1.46) provides the DNA was eluted in 100 mL of preheated (657C) sterile
buffer layer of the waveguides, while the PCR mix water. DNA concentrations were determined by optical
(n = 1.32) provided the vertical side claddings of the density measurements at 260 nm [43] using a spectro-
200-mm-wide waveguides. The SU-8 layer thickness was photometer (Ultrospec 2000, Pharmacia Biotech, Cam-
adjusted to readily accommodate 400 mm OD optical bridge, UK) and the DNA preparations were stored at
fibers (FVP300330370, Polymicro Technologies, L.L.C., 2207C until use. For the testing of the real-time PCR on
USA) in the fiber couplers. The chip was finally packaged the PCR microchip, a C. jejuni DNA template concentra-
on a custom-built aluminum heat sink for passive cooling. tion series (2, 10, 20, 100, and 200 pg/mL) was prepared.
A thin layer of thermal conductivity paste (Dow Corning,
USA) was administered to the backside of the chip to Two primers, namely F2B with sequences 5’-TGG AGG
enhance thermal conduction. GTA ATT TAG ATA TG-3’ and R1B with sequences 5’-CTA
ATA CCT AAA GTT GAA AC-3’ (synthesized by TAG
The electrodes on the chip were connected to the analog Copenhagen, Denmark) were used to amplify a 398-bp
circuitry by two different types of probe pins. Thin (0.37- amplicon of the C. jejuni cadF gene. PCR mixtures (50 mL)
mm diameter) probe pins (SS-30-J-1.3-G, Interconnect contained 0.1 mM (each) dATP, dCTP, dGTP and dTTP
Devices, USA) were used to connect the four thin pads of (TAG Copenhagen), 2 mM MgCl2 (Roche Diagnostics Cor-
the four-point thermometer sensor. Thicker (1.98 mm di- poration, USA), 12.5 nM of each primer (DNA Technology,
ameter) probe pins (S-4-C-5-G, Interconnect Aarhus, Denmark), 61 PCR buffer (Roche Diagnostics

© 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com


5054 Z. Wang et al. Electrophoresis 2006, 27, 5051–5058

Corporation, USA) and 25 U/mL of Taq DNA polymerase is kept short (200 mm), as the measured fluorescent signal
(Sigma, USA), 1 mg/mL non-acetylated BSA (Sigma) and intensity is inversely proportional to the detection dis-
DNA template. Two different DNA binding dyes, SYTOX tance squared. Two such pairs of waveguides were inte-
Orange (ex 547 nm/em 570 nm) and TO-PRO-3 grated at different positions inside the reaction chamber
(ex 642 nm/em 661 nm) (Molecular Probes, USA), were to locally monitor the fluorescence intensity. The design
added into the PCR mix (200 nM each) for the real-time with two pairs of integrated waveguides provides the
PCR detections. ability to monitor two different wavelengths in “multiplex”
real-time PCR applications and can easily be modified to
Of the 50 mL prepared PCR master mix, 25 mL were used
detect fluorescent signals at any desired location inside
for the real-time PCR on the PCR microchip while the rest
the PCR reaction chamber.
was used for a conventional real-time PCR using a com-
mercial real-time PCR machine (Chromo4®, MJ Re- All parts of the integrated optical system in the real time
search, USA) as control. The PCR conditions were: 1cycle PCR microchip are made from SU-8, and the selection of
at 957C for 2 min, followed by 40 cycles of 947C for 15 s, suitable DNA-binding dyes according to the optical
507C for 15 s and 727C for 15 s, and ending with a 2-min properties of SU-8 is therefore a crucial step. SYBR
elongation stage at 727C. The PCR chips were used for a Green I is the most common DNA-binding dye used in
single reaction only. real-time PCR [12, 13, 44], and it is excited by a blue light
source (e.g. an Ar-ion laser with a wavelength of 488 nm).
2.4 Optical system However, SYBR Green I is not a suitable dye for the real-
time PCR microchip described here because SU-8 has
For the real-time PCR measurements on chip, a 60-mW very high light absorption and high fluorescence back-
diode pumped solid-state green laser (535 nm) (DPSSL- ground at low wavelengths (less than 500 nm) [45, 46]. To
60, Viasho, P.R. China) and a 5-mW He-Ne laser (633 nm) avoid this problem, two other DNA-binding dyes, SYTOX
(25-LHR-151-230, Melles Griot, USA) were used to pro- Orange (ex 547 nm/em 570 nm) and TO-PRO-3
vide two different excitation wavelengths. Two different (ex 642 nm/em 661 nm) with longer excitation and emis-
types of PMT from Hamamatsu, Japan, a H5784 (for the sion wavelengths were selected for testing the real-time
535 nm excitation light) and a H5784-01 (for the 633 nm PCR microchip. We chose to test two different dyes be-
excitation light), were used to measure the fluorescent cause these dyes had not been tested in real-time PCR
signals at the two different wavelengths. A 5 mm65 mm before. Tests in conventional real-time PCR showed that
62 mm FGL550S long pass filter (cut from a 50 mm the two dyes were thermally stable and showed very low
650 mm filter slide obtained from Thorlabs, USA) was PCR inhibition (unpublished data).
placed in front of the aperture of the PMT SMA adapter
(E5776-51, Hamamatsu) for filtering of 535 nm excitation In initial experiments, the melting curves were determined
light, while a 5 mm65 mm63 mm LP645 long pass filter by mixing the DNA binding dyes with the DNA fragment
(cut from a 50 mm650 mm filter slide obtained from from a PCR reaction. At the melting point of the DNA
Melles Griot) was placed in front of the aperture of another fragment, where the dsDNA melts to ssDNA, the fluores-
photomulitplier tube (PMT) SMA adapter for filtering of cent signals decrease significantly because the DNA-
633 nm excitation light. To avoid photo bleaching of the binding dyes are only fluorescent when bound to dsDNA.
fluorescence dyes, a custom-modified chopping blade Using the real-time PCR microchip with a temperature
with a 3% duty cycle was placed in front of the lasers. A gradient (from 357C to 957C with 27C/min), the melting
SR540 chopper controller (Stanford Research Systems, curve of the DNA fragment could be determined. The
USA) was used to manipulate the chopper. result of such an experiment with SYTOX Orange-labeled
DNA is shown in Fig. 3. The melting point was determined
by differentiating the registered melting curve, and the
3 Results and discussion melting point (837C) measured on the real-time PCR chip
was the same as measured on the conventional
In this report, a real-time PCR microchip with an inte- Chromo4® real-time PCR thermal cycler. The presented
grated thermal system and a polymer-based optical approach shows good sensitivity for the fluorescence
detection system is presented. The fluorescence from the measurements during the thermal process.
real-time PCR is measured using a pair of waveguides.
The waveguides, one for introducing excitation light, and DNA-binding dye binds to the dsDNA of any PCR prod-
the other for receiving the fluorescent signals, are placed uct. Therefore, specific fluorescence probes (labeled
perpendicular to each other to avoid collection of too ssDNA oligonucleotides) are normally required to distin-
much excitation light. Together they define a detection guish the different products in a multiplex real-time PCR.
point (Fig. 2). The distance between two waveguide ends The costs of the fluorescent DNA-binding dyes are,

© 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com


Electrophoresis 2006, 27, 5051–5058 Miniaturization 5055

Figure 3. Melting curve meas-


urement of the 398-bp amplicon
of Campylobacter jejuni cadF
gene using SYTOX Orange. A
temperature program from 35 to
957C with a ramp of 27C/min
was used. The gradual decrease
in signal in the range from 35 to
807C is due to the temperature-
dependence of the fluores-
cence. The following steep de-
crease depicts the melting pro-
cess of the PCR product. The
melting point was determined
by differentiating the melting
curve (dFL/dTemp), and the
melting point was found to be
837C.

however, much lower than those of the specific fluores-


cent probes, and in some cases, the probes require the
use of an expensive special enzyme. Furthermore, the
requirements for optimization of the PCR protocol are less
stringent when using the DNA-binding dyes. In addition,
by measuring the melting curve after the real-time PCR,
the PCR product can be examined in situ without further
post-PCR analysis (e.g. DNA gel electrophoresis), while it
is not possible when using fluorescence probes. The
DNA-binding dyes were therefore selected for preliminary
testing of the real-time PCR microchip.

During the real-time PCR experiments, two different Figure 4. Typical on-chip real-time PCR data profile
colored fluorescent DNA-binding dyes (SYTOX Orange using TOPRO-3. Due to the thermal dependence of the
and TO-PRO-3) were added into the PCR mixture to fluorescent intensity of the dyes, the three stages of the
monitor the reactions on chip. A typical data profile of the PCR cycle can be clearly distinguished by fluorescence
real-time PCR labeled with TO-PRO-3 is shown in Fig. 4. measurements using the integrated waveguides. The flu-
The three stages of a PCR cycle (15 s denaturation at orescence signal traces correspond perfectly to the
947C, 15 s annealing at 507C and 15 seconds elongation measured temperature profile in the chip.
at 727C) can clearly be distinguished.
Possible surface-induced inhibition is always a critical
In this approach, the temperature deviation during the
issue in microfabricated PCR devices [42, 48–51]. To
temperature switching is less than 0.57C. Such tiny devia-
avoid any inhibition from the SU-8 surface, 1 mg/mL non-
tions are mainly caused by overcompensation through the
acetylated BSA was added to the PCR mix. The BSA
PID control algorithm. A more accurate feedback control
suppresses the interaction of the PCR reagents with the
algorithm may eliminate it. The PCR thermocycling has
SU-8 surface and thus prevents inhibition. The non-
been optimized previously in order to decrease the cycling
acetylated BSA did not add to the background fluores-
time [47]. The achievement of fast cooling (20 6 27C/s) and
cence. Furthermore, the applied dye concentrations in
heating (11 6 17C/s) rates reduced the whole PCR pro-
the PCR mixture are optimized and limited to 200 nM to
cess on-chip to only 30–40 min in comparison to 1.5 h on
avoid any PCR inhibition effects.
a conventional PCR thermocycler (Chromo4®, MJ Re-
search, USA). Therefore, this prototype can be developed By taking the mean fluorescent signal value during the
towards a portable lab-on-a-chip system for rapid elongation period of each thermal cycle, the relative PCR
screening of pathogens in the field. product concentration for each PCR cycle can be deter-

© 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com


5056 Z. Wang et al. Electrophoresis 2006, 27, 5051–5058

mined quantitatively. The results of real-time PCR on chip


using SYTOX Orange to detect the C. jejuni cadF gene
[38] for different DNA template concentrations (from 2 to
200 ng/mL corresponding to about 2.866104 2 2.866107
Campylobacter genome equivalents per chip) are shown
in Fig. 5A. Results of gel electrophoresis for the PCR
products, both on chip and in tube, are shown in Fig. 5B.
The PCR cycle CT for the different DNA template con-
centrations can be clearly determined from the real-time
PCR profile.

The CT value represents the cycle number at which the


PCR product concentration is above the threshold level,
as shown in Eq. (1):

ConcT = Conc06(1 1 e)CT (1)

where ConcT and Conc0 are the concentrations of PCR


product at the threshold level and the initial template
concentration, respectively, and e is the PCR reaction
efficiency. If the efficiency is 100% the concentration will
double for each thermocycle, but in practice, the effi-
ciency will often be somewhat lower. Since the threshold
level is constant, the CT value should be proportional to
the logarithm of the PCR template concentration

CT = mLog(Conc0) 1 b (2)

where m = 2[log(1 1 e)]21 and b = [log(1 1 e)]21log(ConcT).


The efficiency of the PCR reaction can be determined
from the parameter m using Eq. (3)

e = 1021/m 2 1 (3).

Ideally, the final PCR product concentration as measured


by gel electrophoresis should be proportional to the PCR
template concentration, as shown in Eq. (4)

Concproduct ¼ Conc0  ð1 þ eÞCEnd ¼ Conc0  k (4) Figure 5. (A) Results from real-time PCR on chip to
amplify the 398-bp amplicon on the Campylobacter jejuni
where Concproduct is the concentration of the PCR product cadF gene using SYTOX Orange. 0, negative control on
after CEnd thermo cycles and k = (1 1 e)Cend. Equation (4) is chip; 1–5, real-time PCR on chip for detection of the
strictly only valid for cases where the reaction is stopped Campylobacter jejuni DNA template series (2, 10, 20, 100,
in the exponential phase, as in our experiments. For most and 200 ng/mL). (B) Results for gel electrophoresis on an
Agilent Bioanalyzer DNA500 chip. L: DNA marker ladder
PCR, however, the reaction proceeds from exponential to
(15–600 bp); –, negative control in tube; 1, positive con-
linear and finally a plateau phase before the reaction is trol in tube; numbers correspond to the samples men-
terminated and in such cases, the equation is not valid. tioned above.
The raw data has a relatively high noise level due to the
fluorescent background level and the light losses asso-
ciated with SU-8, even with the new dyes. To find the CT and the DNA template concentration (Fig. 6A), and the
value, the raw data (Fig. 5A) were smoothened using an expected linear correlation between the PCR product
eight-point adjacent average function (OriginPro 7.5, data concentration and the DNA template concentration
not shown) and for each graph, a baseline and a line for (Fig. 6B) for both dyes. Linear regression of the CT vs.
the linear amplification range were drawn. The CT value Log(Conc0) plot according to Eq. (2) yields m = 23.95
was then determined as the intercept of these lines. The and b = 30.7 for SYTOX Orange, and m = 24.41 and
results obtained from the real-time PCR microchips b = 34.7 for TO-PRO-3. By using Eq. (3), the reaction
showed the expected logarithmic correlation between CT efficiency can be calculated to be 79 and 69% for SYTOX

© 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com


Electrophoresis 2006, 27, 5051–5058 Miniaturization 5057

with the literature [3, 14]. Furthermore, the dynamic range


of the CT value measurements can be extremely high, e.g.
more than six orders of magnitude [3, 14].

The PCR efficiency on microchip is about 10% lower than


the conventional PCR in tube. This is probably mainly due
to the high surface/volume ratio problem and surface
inhibition [47]. However, the LOD of the PCR microchip is
sufficient for the C. jejuni detection. A better surface/vol-
ume ratio consideration design and good surface coating
that may provide an even higher PCR efficiency of this
microchip are focus points for further development of the
PCR microchip.

4 Concluding remarks
To our knowledge, this is the first time monitoring of real-
time PCR using integrated optical elements in a lab-on-a-
chip system has been demonstrated. All the integrated
polymer optical systems were defined in the same SU-8
layer as the PCR reaction chamber, without any extra
mask step. By using two fluorescent DNA-binding dyes
with suitable excitation and emission spectra (SYTOX
Orange and TO-PRO-3), the progression of the PCR
amplification could be followed efficiently. The measured
CT value on the chip provided more accurate quantitative
information about the initial DNA template concentrations
than the conventional post-PCR analysis by CE. The
integrated optical system described in this study allows
real-time monitoring of the reaction dynamics at any
Figure 6. Two DNA binding dyes (SYTOX Orange and location inside the micro reaction system and can be
TO-PRO-3) with different excitation wavelengths were
integrated into various types of microchips to facilitate
used for detection of the Campylobacter jejuni cadF gene
on the real-time PCR chip. (A) The internal assay (CT) for real-time monitoring for a number of different purposes.
both dyes shows good linear relationships with the loga- Integration of thermal control and polymer waveguides for
rithm of the DNA template concentration series. (B) The real-time PCR is thus an important step towards a port-
external end-point assay results (obtained via DNA gel able microchip system for pathogen detection.
electrophoresis) can only indicate the trend of the DNA
template concentrations. We would like to thank Dr. Klaus B. Mogensen for useful
suggestions for the integrated optical system design. This
research was supported by the Danish Technical Re-
Orange and TO-PRO-3, respectively. Control PCR reac- search Council (STVF) (Grant No. 26-02-0307) and EU
tions in tubes on a conventional PCR thermocycler were STREP project OptoLabCard.
performed in parallel. The results of these controls were:
m = 23.56 and b = 28.1 for SYTOX Orange, and m = 23.59
and b = 32.4 for TO-PRO-3, corresponding to a reaction 5 References
efficiency of 90% and 89% for SYTOX Orange and TO-
[1] Figeys, D., Pinto, D., Anal. Chem. 2000, 72, 330A–335A.
PRO-3, respectively. The parameter b represents the
[2] Wang, Z., El-Ali, J., Nielsen, I. R. P., Mogensen, K. B. et al.,
expected CT value of a sample with 1 pg/mL DNA tem- Lab Chip, 2004, 4, 372–377.
plate. [3] Wilhelm, J., Pingoud, A., ChemBioChem 2003, 4, 1120–1128.
[4] Bernard, P. S., Wittwer, C. T., Clin. Chem. 2000, 46, 147–148.
The CT value measurements (Fig. 6A) provide more accu-
[5] Gibson, U. E., Heid, C. A., Williams, P. M., Genome Res.,
rate information about the initial DNA template concentra- 1996, 6, 995–1001.
tions than the measurements of the final PCR product [6] Kutyavin, I. V., Afonina, I. A., Mills, A., Gorn, V. V. et al., Nucleic
concentration (Fig. 6B). These results are in agreement Acids Res. 2000, 28, 655–661.

© 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com


5058 Z. Wang et al. Electrophoresis 2006, 27, 5051–5058

[7] Tyagi, S., Kramer, F. R., Nat. Biotechnol. 1996, 14, 303–308. [30] Kopp, M. U., de Mello, A. J., Manz, A., Science 1998, 280,
[8] Tyagi, S., Marras, S. A., Kramer, F. R., Nat. Biotechnol. 2000, 1046–1048.
18, 1191–1196. [31] Schneegaß, I., Bräutigam, R., Köhler, J. M., Lab Chip 2001,
[9] Whitcombe, D., Theaker, J., Guy, S. P., Brown, T., Little, S., 1, 42–49.
Nat. Biotechnol. 1999, 17, 804–807. [32] Zhang, Q. T., Wang, W. H., Zhang, H. S., Wang, Y. L., Sens.
[10] Nazarenko, I. A., Bhatnagar, S. K., Hohman, R. J., Nucleic Actuators B 2002, 82, 75–81.
Acids Res. 1997, 25, 2516–2521. [33] Sun, K., Yamaguchi, A., Ishida, Y., Matsuo, S., Misawa, H.,
[11] Nazarenko, I., Lowe, B., Darfler, M., Ikonomi, P. et al., Sens. Actuators B 2002, 84, 283–289.
Nucleic Acids Res. 2002, 30, e37. [34] Wang, W., Li, Z.-X., Luo, R., Lu, S.-H. et al., J. Micromech.
[12] Simpson, D. A., Feeney, S., Boyle, C., Stitt, A. W., Mol. Vis. Microeng. 2005, 15, 1369–1377.
2000, 6, 178–183. [35] Liu, J., Enzelberger, M., Quake, S., Electrophoresis 2002, 23,
1531–1536.
[13] Lekanne Deprez, R. H., Fijnvandraat, A. C., Ruijter, J. M.,
Moorman, A. F., Anal. Biochem. 2002, 307, 63–69. [36] Gulliksen, A., Solli, L., Karlsen, F., Rogne, H. et al., Anal.
Chem. 2004, 76, 9–14.
[14] Wong, M. L., Medrano, J. F., BioTechniques 2005, 39, 75–85.
[37] Gulliksen, A., Solli, L. A., Drese, K. S., Sorensen, O. et al.,
[15] Poser, S., Schulz, T., Dillner, U., Baier, V. et al., Sens. Actua-
Lab Chip 2005, 5, 416–420.
tors A 1997, 62, 672–675.
[38] Bang, D. D., Nielsen, E. M., Scheutz, F., Pedersen, K. et al.,
[16] Daniel, J. H., Iqbal, S., Millington, R. B., Moore, D. F. et al., J. Appl. Microbiol. 2003, 94, 1003–1014.
Sens. Actuators A 1998, 71, 81–88.
[39] Best, E. L., Powell, E. J., Swift, C., Grant, K. A., Frost, J. A.,
[17] Lee, T. M. H., Hsing, I. M., Lao, A. I. K., Carles, M. C., Anal. FEMS Microbiol. Lett. 2003, 229, 237–241.
Chem. 2000, 72, 4242–4247.
[40] Yang, C., Jiang, Y., Huang, K., Zhu, C., Yin, Y., FEMS Immu-
[18] Lin, Y. C., Yang, C. C., Huang, M. Y., Sens. Actuator B- nol. Med. Microbiol. 2003, 38, 265–271.
Chem. 2000, 71, 127–133.
[41] Abu-Halaweh, M., Bates, J., Patel, B. K., Res. Microbiol.
[19] Giordano, B. C., Ferrance, J., Swedberg, S., Huhmer, A. F. 2005, 156, 107–114.
R., Landers, J. P., Anal. Biochem. 2001, 291, 124–132.
[42] El-Ali, J., Perch-Nielsen, I. R., Poulsen, C. R., Bang, D. D. et
[20] Lagally, E. T., Medintz, I., Mathies, R. A., Anal. Chem. 2001, al., Sens. Actuators A 2004, 110, 3–10.
73, 565–570.
[43] Madigan, M. T., Martinku, J. M., Parker, J., Brock Biology of
[21] Nagai, H., Murakami, Y., Yokoyama, K., Tamiya, E., Biosens. Microorganisms, 8th Edn., Prentice-Hall, New Jersey 1997.
Bioelectron. 2001, 16, 1015–1019.
[44] Monis, P. T., Giglio, S., Saint, C. P., Anal. Biochem. 2005,
[22] Yang, J. N., Liu, Y. J., Rauch, C. B., Stevens, R. L. et al., Lab 340, 24–34.
Chip 2002, 2, 179–187.
[45] Marie, R., Schmid, S., Johansson, A., Ejsing, L. et al., Bio-
[23] Yoon, D. S., Lee, Y. S., Lee, Y., Cho, H. J. et al., J. Micro- sens. Bioelectron. 2006, 21, 1327–1332.
mech. Microeng. 2002, 12, 813–823.
[46] Mogensen, K. B., El-Ali, J., Wolff, A., Kutter, J. P., Appl.
[24] Lee, T. M.-H., Carles, M. C., Hsing, I.-M., Lab Chip 2003, 3, Optics 2003, 42, 4072–4079.
100–105. [47] Poulsen, C. R., El-Ali, J., Perch-Nielsen, I. R., Bang, D. D.,
[25] Erill, I., Campoy, S., Rus, J., Fonseca, L. et al., J. Micro- Wolff, A., J. Rapid Methods Autom. Microbiol. 2005, 13,
mech. Microeng. 2004, 14, 1558–1568. 111–126.
[26] Lee, D. S., Park, S. H., Yang, H., Chung, K. H. et al., Lab Chip [48] Shoffner, M. A., Cheng, J., Hvichia, G. E., Kricka, L. J.,
2004, 4, 401–407. Wilding, P., Nucleic Acids Res. 1996, 24, 375–379.
[27] Ke, C., Berney, H., Mathewson, A., Sheehan, M. M., Sens. [49] Shin, Y. S., Cho, K., Lim, S. H., Chung, S. et al., J. Micro-
Actuators B 2004, 102, 308–314. mech. Microeng. 2003, 13, 768–774.
[28] Yan, W., Du, L., Wang, J., Ma, L., Zhu, J., Sens. Actuators B [50] Felbel, J., Bieber, I., Pipper, J., Kohler, J. M., Chem. Eng. J.
2005, 108, 695–699. 2004, 101, 333–338.
[29] Liao, C. S., Lee, G. B., Wu, J. J., Chang, C. C. et al., Biosens. [51] Giordano, B. C., Copeland, E. R., Landers, J. P., Electro-
Bioelectron. 2005, 20, 1341–1348. phoresis 2001, 22, 334–340.

© 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com

You might also like