You are on page 1of 6

Biosensors and Bioelectronics 26 (2010) 1565–1570

Contents lists available at ScienceDirect

Biosensors and Bioelectronics


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

Integration of microfluidic and cantilever technology for biosensing application


in liquid environment夽
Carlo Ricciardi a,∗ , Giancarlo Canavese b , Riccardo Castagna a , Ivan Ferrante a ,
Alessandro Ricci a , Simone Luigi Marasso a , Lucia Napione c , Federico Bussolino c
a
LATEMAR – Politecnico di Torino, Dipartimento di Scienza dei Materiali ed Ingegneria Chimica, Corso Duca degli Abruzzi 24, I-10129 Torino, Italy
b
IIT – Italian Institute of Technology @ POLITO Center for Human Space Robotics, Corso Trieste 21, 10129 Torino, Italy
c
Department of Oncological Sciences, Institute for Cancer Research and Treatment, University of Torino, 10060 Candiolo (TO), Italy

a r t i c l e i n f o a b s t r a c t

Article history: Microcantilever based oscillators have shown the possibility of highly sensitive label-free detection by
Received 14 April 2010 allowing the transduction of a target mass into a resonant frequency shift. Most of such measurements
Received in revised form 9 July 2010 were performed in air or vacuum environment, since immersion in liquid dramatically deteriorates the
Accepted 29 July 2010
mechanical response of the sensor. Besides, the integration of microcantilever detection in a microfluidic
Available online 5 August 2010
platform appears a highly performing technological solution to exploit real time monitoring of biomolec-
ular interactions, while limiting sample handling and promoting portability and automation of routine
Keywords:
diagnostic tests (Point-Of-Care devices). In the present paper, we report on the realization and opti-
Lab-on-Chip
Microcantilever
mization of a microcantilever-based Lab-on-Chip, showing that microplates rather than microbeams
Real-time exhibit largest mass sensitivity in liquid, while pirex rather than polymers represents the best choice for
Protein detection microfluidic channels. Maximum Q factor achieved was 140 (for fifth resonance mode of Pirex prototype),
Angiogenesis as our knowledge the highest value reported in literature for cantilever biosensors resonating in liquid
environment without electronic feedback. Then, we proved the successfully detection of Angiopoietin-1
(a putative marker in tumor progression), showing that the related frequency shifts coming from non-
specific interactions (negative controls) are roughly one order of magnitude lower than typical variations
due to specific protein binding. Furthermore, we monitored the formation of antibody–antigen com-
plex on MC surface in real-time. The proposed tool could be extremely useful for the comprehension of
complex biological systems such as angiogenic machinery and cancer progression.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction nanoscale have recently shown evident performance limits in


terms of analyte density, response time (Nair and Alam, 2006)
Resonance operation method, which aims to quantify the and statistical variability (Gupta et al., 2006) due to their intrinsic
adsorbed target mass thanks to the oscillator frequency shift, seems diffusion-limited regime. Furthermore, such highly sensitive can
to be the most successful application of microcantilever (MC) based be achieved in air or vacuum environment, since immersion in
biosensor. Despite of static operation method (where surface stress liquid would dramatically deteriorate the response of the sensor.
generated in binding of the target molecules to the receptors on one Indeed, minimum detectable mass can be defined as mmin ∝ (m/Q)
MC side cause the beam to deflect), this technique is less affected (Waggoner and Craighead, 2007), where m and Q are the oscilla-
by the thermal drift of beam deflection and stabilization problems tor mass and quality factor, respectively. The latter is defined as
(Shen et al., 2001; Lochon et al., 2006). Q = fr (f−3 dB )−1 , where fr is the resonance frequency and f−3 dB
The progress of microcantilever technology and the need for is the width of the resonance curve at −3 dB from the maxi-
increasing device sensitivity have favored the reduction of sen- mum.
sor dimensions up to the nanoscale. It has been shown that Operation in liquid in fact would be desirable in order to retain
nanocantilevers are able to detect few biomolecules or single biomolecule physiological structure and function, since it has been
viruses (Ilic et al., 2004; Gupta et al., 2004), thus displaying shown that proteins and cell membrane can change their con-
very high mass sensitivity. On the other hand, biosensors at the formation passing from liquid environment to vacuum condition
(Sharma and Kalonia, 2004). Liquid environment is then funda-
mental in those applications where the observation of binding
夽 Finalist for the World Congress on Biosensors 2010 selected Keynote Paper.
and unbinding kinetics are needed as well as for in situ and real-
∗ Corresponding author. Tel.: +39 011 0907383. time measurement. An on-line measurement in liquid can also
E-mail address: carlo.ricciardi@polito.it (C. Ricciardi). reduce false positive and false negative responses, which are an

0956-5663/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.bios.2010.07.114
1566 C. Ricciardi et al. / Biosensors and Bioelectronics 26 (2010) 1565–1570

important drawback in MC-based biosensing (Waggoner and Four main objectives can be identified in the following work: (I)
Craighead, 2007). Moreover, operation in liquid assures the reduc- optimization of sensor geometry (in particular, MC aspect ratio); (II)
tion of the risk of breaking the MCs in transporting them between optimization of microfluidic platform (in particular, best choice for
different environments as well as during the drying procedure. materials of microchannels and interconnections); (III) successfully
Finally, in vacuum conditions, residual salt precipitated from the in liquid detection of Ang-1 (with respect to negative controls); (IV)
buffer solution, employed in cleaning or functionalization steps, observation of antibody–antigen real-time kinetics.
could remain on cantilever surfaces thus invalidating target mass
estimation (Liu et al., 2009). 2. Materials and methods
Recent literature contains some interesting studies about can-
tilevers vibrating in liquid performed under static fluid conditions 2.1. Reagents
(Campbell and Mutharasan, 2005a; Park et al., 2005). Campbell and
Mutharasan (2005b) performed measurements using millimeter- 3-Aminopropyltriehoxysilane (APTES, anhydrous, 99% Aldrich),
sized cantilever partially immersed in a static liquid; this approach glutaraldehyde (GA, 25%, v/v, water solution) and toluene
has the disadvantage to register a decrease of volume during the (anhydrous, 99.8% Aldrich) were used without any further purifi-
experiment due to the evaporation and consequently it requires a cation. Sulphuric acid (95–97%, w/w) and hydrogen peroxide
more complex calibration. It is worth to note that almost all the (30%, w/w) were also purchased from Sigma–Aldrich. Ortho-
experiments concerning cantilevers immersed in liquid are carried boric acid and sodium chloride used to prepare borate buffer
out in commercial or home-made fluid cell (Arntz et al., 2003; Braun were ACS reagents (assay ≥ 99.5%) and were obtained from
et al., 2005; Ghatkesar et al., 2008; Moulin et al., 2000; Vančura et Sigma–Aldrich. Recombinant Protein G, purified from Streptococ-
al., 2005) with volume from tens of microliters to 2 mL, in which cus, was from PIERCE. Ang-1 was obtained from R&D Systems,
the device is placed and sealed through o-ring or rubber membrane. while the anti-Ang-1 (A0604) mAb was from SIGMA. Dulbecco’s
Maraldo et al. (2007a, 2007b) and Campbell et al. (2007) obtained Phosphate-Buffered Saline (PBS; BE17-512F) was procured from
great results in biomolecules mass detection at different concentra- BioWhittaker/Cambrex.
tions by piezoelectric-excited millimeter-sized cantilever (PEMC).
Since the cantilevers are made individually and manually, a low 2.2. LOC fabrication
reproducibility and repeatability in fabrication process is obtained.
In addition their design is difficult to be integrated in standard low The cantilever chips were fabricated starting from Silicon On
cost production both for what concerns the materials and the pro- Insulator (SOI) wafers with the following layers and thicknesses:
cesses. The large dimension of the PEMC forces to design a wide polished silicon device layer of 7 ± 0.5 ␮m, silicon handle layer of
housing flow cell and the fabrication processes are not easy to 450 ± 5 ␮m, buried and backside oxide layers 1.5 ± 0.1 ␮m. The pro-
integrate in Lab-on-Chip (LOC) technology. cess flow is composed of the following steps: photolithography on
Very few works report a fluidic circuit integrated on microcan- the back side, wet etching in Buffered Oxide Etch (BOE) solution for
tilever sensor device (Park et al., 2008; Thaysen et al., 2001). Aubin the patterning of the mask layer on the back side, front side pro-
et al. (2005) presented a design of integration between resonator tection with a polymeric coating (Protek B2 from Brewer Science),
sensor array and microfluidic channel using micromachining pro- wet etching in KOH solution, sample cleaning, photolithography
cesses. Since the dimensions of their resonators are about 10 ␮m, on the front side of the membrane, Reactive Ion Etching (RIE) of
during the measurements, the microfluidic channels hosting the silicon, removal of the buried oxide layer in BOE and cleaning
sensors are pumped down to a pressure where damping effects by piranha solution (Canavese et al., 2007). The complete process
become negligible. This last solution appears to be expensive and flow, together with additional MC fabrication details, is reported in
does not promote integration and device portability. Supplementary material.
It is clear, therefore, that MC biosensors would benefit in Three different materials (SU-8, PDMS, Pirex) and relative
case of integration with microfluidic technology; advantages process fabrication were chosen for three different microfluidic
such as reduction of volume reagents and time assay, in situ prototypes of LOC for the integration with microcantilever sen-
measurements and high level of automation are evident. More- sors. All layouts have at least two wells (roughly 2 ␮L in volume,
over, diagnostic biosensing protocols, as well as single-molecule each), in order to perform different biodetection experiments
detection, could express their practical potentiality only with simultaneously. Polydimethylsiloxane (PDMS) was selected for the
microfluidic assembly, where total fluidic volumes is minimized, fabrication of the microfluidic interconnection for its peculiar char-
recirculation procedure can be created to achieve enhanced acteristics as biocompatibility, transparency, cost-efficiency, etc.
binding probabilities, and the environment to which the sensor is A particular interconnection layout requiring no external clamp-
exposed can be actively controlled. ing or gluing was studied. For all the prototypes were adopted the
In this paper, we report on the realization and optimization same PDMS microfluidic reversible interconnections designed and
of a MC-based LOC to perform in liquid real-time detection of fabricated as described in a previous study (Quaglio et al., 2008).
specific proteins. We chose Angiopoietin-1 (Ang-1) as our target A picture and a 3D sketch of the device, together with impor-
molecule representative of the wide range of angiogenic factors, tant parts of the experimental set-up are reported in Fig. 1a and b,
whose expression level is largely investigated in different tumors. respectively.
Angiopoietin-1 and -2 are oligomeric-secreted glycoprotein ligands Processes flow description as well as fabrication details are
of the receptor tyrosine kinase Tie-2 (Davis et al., 1996). Even if reported in Supplementary material.
it is well established that the Ang–Tie-2 pathway is involved in
tumor angiogenesis, the exact effects of angiopoietins on tumor 2.3. Experimental set-up and data analysis
angiogenesis are under debate (Metheny-Barlow and Li, 2003).
Nevertheless experimental and clinical studies have demonstrated Cantilever vibrational characteristics were measured using an
that increased expression of Ang-1 and -2 promotes or inhibits apparatus in which a function generator (HP 33120A) produced
tumor angiogenesis (Yu, 2005), suggesting that Ang-1 is a pro- a sinusoidal signal that was amplified and sent to a piezoelectric
angiogenic factor that promotes endothelial cell survival and tumor actuator (PI Ceramic). Cantilever resonance curves were monitored
angiogenesis, especially in the presence of vascular endothelial using an optical lever read-out (laser diode and PSD by Hama-
growth factor. matsu): the current output of the PSD was amplified and converted
C. Ricciardi et al. / Biosensors and Bioelectronics 26 (2010) 1565–1570 1567

tion for few data can lead to an underestimation of the uncertainty).


Finally, we use the “weighted average” method to have the best esti-
mation of the true value f/f of all the MCs, as reported elsewhere
(Ricciardi et al., 2010).

2.4. MC functionalization and binding assays

Silicon MCs were soaked into piranha solution (70% H2 SO4 :30%
H2 O2 ) for 15 min, rinsed with doubly distilled water and dried in
a stream of nitrogen; after that, a silicon oxide film was grown
on them by thermal oxidation at 1100 ◦ C in O2 atmosphere for
3 h. Freshly cleaned MC were then incubated with 1% (v/v) solu-
tion of APTES in toluene at 60 ◦ C for 10 min. Silane-coated MCs
were rinsed several times with toluene and dried with nitrogen.
Freshly silanized-MCs were then incubated in a 1% GA solution
(in borate buffer 0.1 M, pH 8.8) for 1 h using an orbital shaker at
100 rpm. After 15 min of incubation, 300 ␮L of a solution of sodi-
umcyanoboro hydride (5 M) in NaOH have been added to reduce
the imine (–N CH–) formed from the reaction between –NH2 and
–CHO moieties. After incubation, MCs have been rinsed several
times with doubly distilled water and dried in a stream of nitro-
gen (Ricciardi et al., 2010). Then, 60 ␮L of protein G solution (PtG,
0.05 mg/mL) in PBS was pumped at 0.5 ␮L/min through microfluidic
channels on functionalized cantilever for 2 h at room tempera-
ture. Afterward, Tween 20 (0.05% in PBS) was pumped for 30 min
to remove non-specifically bound proteins. After protein G bind-
ing, 30 ␮L of anti-Ang-1 mAb (A0604) solution in PBS (0.05 mg/mL)
were pumped on cantilever surface and the incubation performed
Fig. 1. Microcantilever-based LOC: (a) picture and (b) 3D sketch. Most important for 2 h at room temperature. After this, channels and cantilevers
parts of the device are labeled as: (1) PDMS interconnections; (2) cantilever chip;
were washed with PBS. Antigen recognition was carried out by
(3) microfluidic platform (Pirex, SU-8 or PDMS); (4) piezo disk; (5) Peltier cell; (6)
heat sink.
flowing 30 ␮L of Ang-1 solution in PBS (25 ␮g/mL) at controlled
temperature 23 ± 0.2 ◦ C at a flow rate of 0.5 ␮L/min.

into a voltage output, sent to a lock-in amplifier (EG&G 7260) for


3. Theory
signal extraction and filtering, and stored to a personal computer.
To properly mechanically lock the chip to the piezo disc and to
According to Euler–Bernoulli theory, the nth mode resonance
improve the heat transfer between Peltier element (PF-127-14-
frequency of a rectangular cantilever beam in vacuum fnv is given
15, 40 mm × 40 mm × 3.9 mm, 58 W, SuperCool) and chip, thermal
by:
glue was employed (Thermal Bonding System by Electrolube Der-

byshire, UK). The temperature was set by a temperature controller 2 E t
(LDT-5948 by ILX Lightwave) and the feedback signal necessary fnv = n (2)
2 12 l2
to regulate the temperature was obtained interfacing the Propor-
tional Integral Derivative (PID) controller to a thermistor placed in where E and  are respectively the elastic modulus and the
contact to the upper surface of the piezo element: PID coefficients density of the cantilever material, while t and l are the beam
were obtained using the AutoTune function of the temperature con- thickness and length. The constants n are dimensionless parame-
troller. The temperature was stabilized in the range of 0.1 ◦ C. Liquids ters depending on boundary conditions. For rectangular cantilever
are injected inside the LOC by means of a syringe pump (by Harvard beam, the numerical values are: 1 = 1.875, 2 = 4.694, 3 = 7.85, . . .,
Apparatus) that allows for a fine setting of pumping rate. n = (n − 0.5) when n > 5 (Lochon et al., 2005).
The measurement procedure as well as the fitting of data with a When such a structure oscillates in a viscous fluid, inertial and
Lorentzian curve were performed by a software in LABVIEW® envi- viscous forces act against its motion so that resonant frequency
ronment. The resonant frequency f and the quality factor Q of each and quality factor Q are significantly affected. In the case of an
single curve were extracted from fitting the measured amplitude unbounded fluid environment, an analytical model, holding for low
of vibration as function of the excitation frequency, A(f) as follows: mode numbers, is available for estimating the fluid induced Q fac-
 tors and resonance frequencies, in the case in which the magnitude
Amax 4Q 2 − 1 of the dissipative effects is small, i.e. Q  1 (Sader, 1998). Defining
A(f ) = b1 + b2 f + (1)
2Q 2 2 2 f ,  and w respectively the fluid density, the fluid viscosity and
(1 − (f 2 /fr2 )) + (f /Qfr )
the beam width, the nth mode resonance frequency in fluid can be
where Amax is the maximum amplitude of vibration and b1 and b2 written as (Sader, 1998):
are baseline fitting coefficients. fnv
f
Since we used MC with slightly different geometrical dimen- fn =  (3)
1 + (f /4t)(c1 ı + c2 w)
sions (due to inherent tolerances in the fabrication process) and
three resonance modes are usually detected, bioexperiments are being:
compared in terms of relative frequency deviation f/f rather than 
absolute frequency shift. For each MC, we calculate the arithmetic 2
mean of relative frequency deviation f/f over the modes and the ı= (4)
f
2f fn
uncertainty as half-deviation of the modes (using standard devia-
1568 C. Ricciardi et al. / Biosensors and Bioelectronics 26 (2010) 1565–1570

Fig. 2. Histogram of Q factor of the first three normal modes of microcantilevers


with different aspect ratios (3-2-1.5) resonating in water.
Fig. 3. Histogram of Q factor of the first five normal modes of microcantilevers
(AR = 1.5) with different microfluidic channel materials: Pirex, SU-8, PDMS.

The parameter ı represents the unsteady viscous layer thickness,


i.e. thickness of the fluid layer, sticking to the cantilever surfaces,
quality factors (extrapolated from fit introduced in Eq. (1)) of the
where viscous forces are significant. In the frequency range of
first three normal modes (m1 , m2 and m3 ) in water environment as
1–200 kHz, values of viscous layer in water environment are in the
a function of the three different aspect ratios. It arises that Q factor
range of 1–10 ␮m.
can be easily tuned both employing higher modes and reducing the
For what concerns mode Q factors, one has:
cantilever aspect ratio, from a minimum of 5 to a maximum of 24,
(4wt/f ) + c1 ıw + c2 w2 a value comparable with most of similar literature works. There-
Qn = (5) fore, although being quite counterintuitive, minimum detectable
c3 ı2 + c4 ıw
mass of microplates (structures with l∼w) increments with oscil-
Values of coefficients c1–4 can be found in Maali et al. (2005) as: lator width, as theoretically demonstrated for beams (structures
c1 = 3.7997, c2 = 1.0553, c3 = 2.7374, c4 = 3.8018 (6) with l  w) in Section 3. On the other hand, the increment of Q
with higher modes is simply due to the reduction of viscous layer
Since ı only slightly depends on w (as one can verify by simply iter- (Eq. (4)), and thus dumping effects, with frequency increment.
atively solving Eqs. (3) and (4) for a given value of fnv ), a nearly linear The second objective of the work mainly concerns with find-
increase of Q with w is obtained from Eq. (5). Therefore, wider struc- ing the best material to be employed for the microfluidic platform.
tures are expected to exhibit narrower curves (and thus enhanced Such material should generally be compatible with biological pro-
mass detection limits) in fluid, a behavior that commonly seems tocols and LOC technology, but, in particular, it should also exhibit
a little surprising and counterintuitive. In other words, this means suitable mechanical and vibrational characteristics. Due to our
that a reduction of microcantilever planar aspect ratio (AR) = l/w, peculiar design, the microfluidic platform is also responsible for
while keeping constant l and t, is in favor of an increase of Q, a fact the propagation of mechanical vibration from the piezo actuator
that also finds confirmation in experiments (Vančura et al., 2008). to MC (Fig. 1b). Therefore, we fabricated three different prototypes
It is worth to point out anyway that the above treatment is in order to investigate the influence on the sensor performance
strictly valid for beams, i.e. structures for which the condition of the following materials: Pirex, SU-8, PDMS (for details con-
l  w  t holds (typical values are l ≈ 10w ≈ 100t). Therefore, in cerning device fabrication, please refer to Supplementary text
case of low AR structures such as cantilever microplates (i.e. when and figures). The Si MCs have nominally the same dimensions
w∼l) an experimental validation is still needed. (900 ␮m × 600 ␮m × 6 ␮m), except for unavoidable fabrication tol-
erances and SOI substrates thickness uncertainties. Fig. 3 shows
4. Results and discussion the experimental quality factors (extrapolated from fit introduced
in Eq. (1)) of the first five normal modes (m1 , m2 , m3 , m4 and m5 )
4.1. Design optimization in water environment for the three prototypes. It can be clearly
seen that the presence of a soft polymeric layer (SU-8 or PDMS)
First objective of the work is the optimization of device geom- between the piezo actuator and the cantilever seriously affects
etry, in particular MC aspect ratio (AR). According to above the oscillator response. While at low frequency (first three modes,
guidelines, we designed and realized rather wide microstruc- f ≤ 35 kHz) Q factors are similar, at high frequencies (fourth and fifth
tures to achieve relative high Q resonators. While thickness and mode, f ≥ 75 kHz) the damping effects of the polymeric platform are
length were respectively fixed to 6 ␮m and 900 ␮m, three val- so relevant that the resonance peaks completely flatten for PDMS
ues of MC width were tested: 300 ␮m (i.e. AR = 3), 450 ␮m (i.e. chips, drastically weaken for SU-8 chips (which has a larger stiff-
AR = 2), 600 ␮m (i.e. AR = 1.5). Experimentally, first five flexural res- ness respect to PDMS). Raw data and fittings related to Fig. 3 are
onance modes of such structures when vibrating in water are in the reported in Supplementary Material.
following ranges: m1 ∼ = 1–2 kHz, m2 ∼= 8–12 kHz, m3 ∼= 24–35 kHz, Maximum Q factor achieved was 140 (for m5 of Pirex proto-
m4 ∼= 50–75 kHz, m5 ∼ = 90–150 kHz (differences in resonance fre- type), as our knowledge the highest value reported in literature for
quencies are due to unavoidable fabrication tolerances and SOI MC biosensors resonating in liquid environment without electronic
substrates thickness uncertainties). Fig. 2 shows the experimental feedback.
C. Ricciardi et al. / Biosensors and Bioelectronics 26 (2010) 1565–1570 1569

Table 1
Measured frequency shifts and relative frequency changes for Ang-1 at 25 ␮g/mL,
negative control 1 and 2. See text for further details.

Chip Bioexperiment Relative frequency


change, f/f (×10−2 )

FC1 Ang-1 25 ␮g/mL −1.12 ± 0.72


Negative control 1 0.17 ± 0.20

FC2 Ang-1 25 ␮g/mL −1.52 ± 0.94


Negative control 2 −0.10 ± 1.05

FC3 Ang-1 25 ␮g/mL −2.80 ± 1.54


Negative control 2 0.47 ± 0.53

FC4 Ang-1 25 ␮g/mL −6.10 ± 3.77


N/A N/A

Fig. 4. Histogram of average relative frequency deviation for Ang-1 specific binding,
negative control 1 (PBS on Ab-coated MC) and negative control 2 (Ang-1 on Ab-
uncoated MC).

4.2. Biosensing experiments

The high quality factor in a viscous liquid environment obtained


thanks to the integration of new cantilever design with microflu-
idic technology and the use of proper materials entail to perform in
situ real time monitoring of bio-molecular interactions by charac-
terizing resonance frequency shift induced by the specific binding
occurred on the Ab-immobilized cantilever surface. Since highest
Q values were obtained with the pirex-silicon chip, the third device
design has been chosen for biosensing experiments of Ang-1 detec-
tion. MC dimensions were increased to 1200 ␮m × 800 ␮m × 7 ␮m,
while keeping AR = 1.5, to further enhance resonator quality fac-
tor. As reported in Section 2, the biodesign is composed by 2 Fig. 5. Real-time monitoring of antigen–antibody hybridization compared to nega-
functionalization steps (APTES and GA), and 3 biomolecule bind- tive control experiment (PBS on Ab-coated MC).
ing steps (Protein G, anti-Ang-1 Ab, Ang-1); previous experiments
in vacuum environment have shown that such biodesign exhibits
optimal specificity and fine precision, with a ratio of Ang-1/Ab sur- (−0.17 ± 0.20) × 10−2 , (f /f )blank2 = (−0.36 ± 0.470) × 10−2 . As
face density of 2.06 ± 0.05 (Ricciardi et al., 2010). We fabricated can be noticed, the average relative frequency shift of negative con-
and functionalized five chips (FC1-5) that present two indepen- trol experiments resulted roughly one order of magnitude lower
dent cantilevers and wells on the same LOC platform (Fig. 1): the than successfully Ang-1 detection. As could be expected the non-
first resonator is used for the Ang-1 specific recognition, while the specific protein–protein interaction between Ang-1 and protein G
second is used as a negative control. We considered two type of neg- coated MC (negative control 2, Ang-1 on Ab-uncoated MC) is a
ative control (also referred as “blank”): (1) PBS solution without the larger interferon respect to standard blank experiments (negative
target analyte flowing on Ab-coated MC, and (2) Ang-1 solution in control 1, PBS on Ab-coated MC). Calculating the ratio of Ang-1/Ab
PBS (25 ␮g/mL) flowing on Ab-uncoated MC (coated with Protein surface density, we obtain 5.2 ± 1.4, a value characterized by a con-
G). Such negative controls are fundamental to determine the influ- siderably larger relative uncertainty, as expected for measurements
ence of non-specific adsorption/desorption on resonance curves, in liquid respect to previously reported results in vacuum.
therefore setting up the experimental limit of detection of the sys- Last fabricated and functionalized chip (FC5) was used for the
tem. All the solutions were delivered at a flow rate of 0.5 ␮L/min at real-time monitoring of antibody–antigen biomolecular interac-
controlled temperature 23 ± 0.2 ◦ C. tions, as reported in Fig. 5. After few minutes, probably needed
Since we used MC with slightly different geometrical dimen- for the antigen to reach a minimal detectable concentration on
sions (due to inherent tolerances in the fabrication process) and MC surface, the (normalized) resonant frequency of the Ab-coated
three resonance modes are usually detected, bioexperiments are MC clearly decreases and quickly reaches saturation. On the con-
compared in terms of relative frequency deviation f/f rather than trary, the signal coming from negative control (PBS on Ab-coated
absolute frequency shift. For each MC, we calculate the arith- MC) just exhibits negligible fluctuations. Similar measurements
metic mean of relative frequency deviation f/f over the modes at different concentrations are planned to deeply investigate the
and the uncertainty as half-deviation of the modes. Finally, the Ab/Ang-1 binding model and kinetics, as well as to better compare
“weighted average” method is applied to have the best estimation the stoichiometry of such interaction in liquid environment with
of the true value f /f of all the MCs (Ricciardi et al., 2010). Fig. 4 previously reported results in vacuum.
shows the histograms of average relative frequency deviation for
Ang-1 specific binding, negative control 1 (PBS on Ab-coated MC) 5. Conclusions
and negative control 2 (Ang-1 on Ab-uncoated MC), while data
for each MC are reported in Table 1. The calculated values are, We report on the realization and optimization of a
respectively: (f /f )Ang-1 = (−1.59 ± 0.52) × 10−2 , (f /f )blank1 = microcantilever-based LOC, showing that microplates rather
1570 C. Ricciardi et al. / Biosensors and Bioelectronics 26 (2010) 1565–1570

than microbeams exhibit largest mass sensitivity in liquid, while Campbell, G.A., Uknalis, J., Tu, S.I., Mutharasan, R., 2007. Biosens. Bioelectron. 22,
pirex rather than polymers represents the best choice for microflu- 1296–1302.
Canavese, G., Marasso, S.L., Quaglio, M., Cocuzza, M., Ricciardi, C., Pirri, C.F., 2007. J.
idic channels. Maximum Q factor achieved was 140 (for fifth Micromech. Microeng. 17, 1387–1393.
resonance mode of Pirex prototype), as our knowledge the highest Davis, S., Aldrich, T.H., Jones, P.F., Acheson, A., Compton, D.L., Jain, V., Ryan, T.E.,
value reported in literature for cantilever biosensors resonating in Bruno, J., Radziejewski, C., Maisonpierre, P.C., Yancopoulos, G.D., 1996. Cell 87,
1161–1169.
liquid environment without electronic feedback. Then, we proved Ghatkesar, M.K., Braun, T., Barwich, V., Ramseyer, J.P., Gerber, C., Hegner, M., Lang,
the successfully detection of Angiopoietin-1 (a putative marker H.P., 2008. Appl. Phys. Lett. 92, 043106.
in tumor progression), showing that the related frequency shifts Gupta, A., Akin, D., Bashir, R., 2004. Appl. Phys. Lett. 84, 1976–1978.
Gupta, A.K., Nair, P.R., Akin, D., Ladisch, M.R., Broyles, S., Alam, M.A., Bashir, R., 2006.
coming from non-specific interactions (negative controls) are Proc. Natl. Acad. Sci. U.S.A. 103, 13362–13367.
roughly one order of magnitude lower than typical variations due Ilic, B., Craighead, H.G., Krylov, S., Senaratne, W., Ober, C., Neuzil, P., 2004. J. Appl.
to specific protein binding. Furthermore, we monitored the forma- Phys. 95, 3694–3703.
Liu, Y., Li, X., Zhang, Z., Zuo, G., Cheng, Z., Yu, H., 2009. Biomed. Microdev. 11, 183–
tion of antibody–antigen complex on MC surface in real-time. The
191.
proposed tool could be extremely useful for the comprehension Lochon, F., Dufour, I., Rebière, D., 2005. Sens. Actuators B 108, 979–985.
of complex biological systems such as angiogenic machinery and Lochon, F., Dufour, I., Rebière, D., 2006. Sens. Actuators B 118, 292–296.
cancer progression. Maali, A., Hurth, C., Boisgard, R., Jai, C., Cohen-Bouhacina, T., Aimé, J.P., 2005. J. Appl.
Phys. 97, 074907.
Maraldo, D., Rijal, K., Campbell, G., Mutharasan, R., 2007a. Anal. Chem. 79,
Acknowledgements 2762–2770.
Maraldo, D., Garcia, F.U., Mutharasan, R., 2007b. Anal. Chem. 79, 7683–7690.
Metheny-Barlow, L.J., Li, L.Y., 2003. Cell Res. 13, 309–317.
The authors are thankful for the financial support of MIUR Moulin, A.M., O’Shea, S.J., Welland, M.E., 2000. Ultramicroscopy 82, 23–31.
(FIRB2003-LATEMAR and PRIN2007 grant) and Regione Piemonte Nair, P.R., Alam, M.A., 2006. Appl. Phys. Lett. 88, 233120.
(Converging Technologies, NAMATECH grant). Park, J.H., Know, T.Y., Yoon, D.S., Kim, H, Kim, T.S., 2005. Adv. Funct. Mater. 15,
2021–2028.
Park, K., Jang, J., Irimia, D., Sturgis, J., Lee, J., Robinson, P., Tonerd, M., Bashir, R., 2008.
Appendix A. Supplementary data Lab Chip 8, 1034–1041.
Quaglio, M., Canavese, G., Giuri, E., Marasso, L.S., Perrone, D., Cocuzza, M., Pirri, C.F.,
2008. J. Micromech. Microeng. 18, 055012.
Supplementary data associated with this article can be found, in Ricciardi, C., Fiorilli, S., Bianco, S., Canavese, G., astagna, R., Ferrante, I., Digrego-
the online version, at doi:10.1016/j.bios.2010.07.114. rio, G., Marasso, L.S., Napione, L., Bussolino, F., 2010. Biosens. Bioelectron. 25,
1193–1198.
Sader, J.E., 1998. J. Appl. Phys. 84, 64–76.
References Sharma, V.K., Kalonia, D.S., 2004. AAPS PharmSciTech 17 5 (1), E10.
Shen, F., Lu, P., O’Shea, S.J., Lee, K.H., Ng, T.Y., 2001. Sens. Actuators A 15, 17–23.
Arntz, Y., Seelig, J.D., Lang, H.P., Zhang, J., Hunziker, P., Ramseyer, J.P., Meyer, E., Thaysen, J., Marie, R., Boisen, A., 2001. Proc. IEEE, 401–404 (MEMS, 2001).
Hegner, M., Gerber, C., 2003. Nanotechnology 14, 86–90. Vančura, C., Lichtenberg, J., Hierlemann, A., Josse, F., 2005. Appl. Phys. Lett. 87,
Aubin, K.L., Park, S.M., Huang, J., Craighead, H.G., Ilic, B.R., 2005. Sensors (2005 IEEE). 162510.
Braun, T., Barwich, V., Ghatkesar, M.K., Bredekamp, A.H., Gerber, C., Hegner, M., Lang, Vančura, C., Dufour, I., Heinrich, S.M., Josse, F., Hierlemann, A., 2008. Sens. Actuators
H.P., 2005. Phys. Rev. E 72, 031907. A 141, 43–51.
Campbell, G.A., Mutharasan, R., 2005a. Sens. Actuators A 122, 326–334. Waggoner, P.S., Craighead, H.G., 2007. Lab Chip 7, 1238–1255.
Campbell, G.A., Mutharasan, R., 2005b. Biosens. Bioelectron. 21, 462–473. Yu, Q., 2005. Future Oncol. 1, 475–484.

You might also like