Вы находитесь на странице: 1из 21

Anal Bioanal Chem (2013) 405:57855805

DOI 10.1007/s00216-013-7013-z

REVIEW

Inkjet printed (bio)chemical sensing devices


Nobutoshi Komuro & Shunsuke Takaki & Koji Suzuki & Daniel Citterio

Received: 7 March 2013 / Revised: 19 April 2013 / Accepted: 23 April 2013 / Published online: 16 May 2013
# Springer-Verlag Berlin Heidelberg 2013

Abstract Inkjet printing has evolved from an office printing application to become an important tool in industrial
mass fabrication. In parallel, this technology is increasingly
used in research laboratories around the world for the fabrication of entire (bio)chemical sensing devices or single
functional elements of such devices. Regularly stated characteristics of inkjet printing making it attractive to replace an
alternative material deposition method are low cost, simplicity, high resolution, speed, reproducibility, flexibility, noncontact, and low amount of waste generated. With this
review, we give an overview over areas of (bio)chemical
sensing device development profiting from inkjet printing
applications. A variety of printable functional sensor elements are introduced by examples, and the advantages and
challenges of the inkjet method are pointed out. It is demonstrated that inkjet printing is already a routine tool for the
fabrication of some (bio)chemical sensing devices, but also
that novel applications are being continuously developed.
Finally, some inherent limitations of the method and challenges for the further exploitation of this technology are
pointed out.
Keywords Inkjet printing . Screen printing . Electrodes .
Conducting polymers . Microfluidic paper-based analytical
devices
Abbreviations
DBSA
Dodecylbenzenesulfonic acid
DOD
Drop-on-demand
Nobutoshi Komuro and Shunsuke Takaki have equally contributed to
this work.
N. Komuro : S. Takaki : K. Suzuki : D. Citterio (*)
Department of Applied Chemistry, Keio University,
3-14-1 Hiyoshi, Kohoku-ku,
223-8522 Yokohama, Japan
e-mail: citterio@applc.keio.ac.jp

HRP
LbL
LED
LOD
ODF
PANI
PEDOT-PSS
SERS

Horseradish peroxidase
Layer-by-layer
Light emitting diode
Limit of detection
Oligodeoxyfluoroside
Polyaniline
Poly(3,4-ethylene
dioxythiophene)-poly(styrenesulfonate)
Surface enhanced Raman
scattering

Introduction
The application of inkjet printing technology is by far no
longer limited to the printing of text or graphical data onto
paper, films, or three-dimensional objects. The probably
earliest report on the application of inkjet printing for
(bio)sensor fabrication was published in 1988 [1]. The authors of that work looked at inkjet technology to tackle the
challenge of selectively depositing an active sensing layer
onto a small area of a field effect transistor. Over the past
years, inkjet printing has evolved into a general industrial
fabrication tool for depositing controlled small amounts of
liquids onto a user-selected, well-defined area. Inkjet printing technology has grown up to industrial-scale mass production, where it is, for example, nowadays used in the
fabrication of color filters for light-emitting diodes and
full-color high-resolution flat panel displays [2]. The application of inkjet printing technology for material deposition
and device fabrication has been regularly reviewed in the
past [36]. Promoted by the rapid expansion of application
areas and the technical progress in printing equipment,
inkjet printing technology has relatively early found its role
as a tool supporting the mass production of (bio)analytical
devices. Probably among the earliest routine applications

5786

was the use alternative to contact printing arrayers for the


deposition of DNA samples in the high throughput fabrication of DNA analysis chips [7]. Since then, the prospects of
simple laboratory scale prototyping and of mass fabrication
at comparably low costs have remained driving forces for
the further expansion of inkjet printing technology into the
field of (bio)analytical chemistry, including the fabrication
of (bio)chemical sensing devices.
In a previous review, Gonzalez-Macia et al. have given
an overview over advanced printing and deposition methodologies for the fabrication of biosensors and biodevices,
where inkjet technology plays an essential role [8]. Delaney
et al. took a closer look at inkjet printing applied to protein
deposition [9], and Kukkola et al. published a short overview of inkjet printed gas sensors [10]. With this critical
review, we particularly address the (bio)chemical sensor
community by looking at the already widespread and further
increasing application of inkjet printing in the context of the
fabrication of (bio)chemical sensing devices in general. The
terms sensor and sensing devices are used in a broad
sense to include also analytical systems that are probably not
regarded as sensors under a very strict definition. Based on
selected, nonexhaustive examples, it is shown that inkjet
printing is already a routine tool for the fabrication of some
(bio)chemical sensing devices, and that novel applications
are being continuously developed. With the prospects of
simple laboratory scale prototyping and low cost mass fabrication in mind, discussion in this work is limited to the
application of currently commonly available inkjet printing equipment. This is to be understood as covering a
variety of inkjet printing devices reaching from simple consumer desktop printers to the more versatile material printing systems for research use.
In the fabrication of (bio)chemical sensing devices, inkjet
printing systems are in most situations used as one fabrication tool among others to perform just a single specific
material deposition task in the multi-step process leading
to a complete sensing device. Very often this deposition step
concerns the formation of an electrically conducting trace
(e.g., electrode, electric contact) or the active sensing layer
(e.g., polymer film, enzyme or antibody spot, colorimetric
reagent, etc.). The first application field strongly profits
from developments achieved in printed electronics [11]. In
particular, in the fabrication of metallic planar electrodes
(Ag, Au) for electrochemical sensors, inkjet printing can
already be regarded as a routine technology. This is best
reflected by the fact that various ready-to-use inkjet
printable metal nanopaste inks are commercially available. In more recent approaches, inkjet printing is applied for the microfluidic patterning of paper sensor
substrates. Finally, attempts are made to fabricate entire sensing devices (all inkjet printed devices) relying on inkjet
printing technology alone.

N. Komuro et al.

We try to provide some answers, in particular to the


question why inkjet printing has been selected as the method
of choice among the available alternative technologies.

General comparison of inkjet printing and alternative


technologies commonly applied in the fabrication
of (bio)chemical sensing devices
The principal function performed by an inkjet printing system
used as a fabrication tool is the controlled deposition of small
droplets of liquids (ink) onto a substrate (Fig. 1). After the
evaporation of the solvent, the nonvolatile ink components are
left over as the deposited target material. When evaluating the
usefulness of inkjet printing technology in the context of the
fabrication of (bio)chemical sensing devices, this material
deposition function of the inkjet printer can be compared
with other methods fulfilling an equivalent purpose. Of particular relevance are (1) drop casting (manual pipetting), (2)
contact printing, (3) pneumatic dispensing, (4) screen printing,
(5) spin coating, and (6) photolithography, among others.
From a merely process mechanics point of view, drop
casting, contact printing, and pneumatic dispensing are most
closely related to inkjet printing technology. All methods
transfer a controlled amount of liquid from a reservoir (e.g.,
pipette, ink tank, microplate, etc.) to user-selected spots on a
substrate. They all allow flexible spatial control of the material
deposition process. However, when it comes to achievable
resolution, there are significant differences. Both drop casting
and pneumatic dispensing are inferior to inkjet printing, mainly
because of the differences in dispensed volumes. The volume
of the liquid droplet determines to a large extent the resolution
of the deposited feature. Reproducible droplet formation in the
picoliter volume order is nowadays routinely possible with
almost any type of inkjet printing device (Fig. 2), whereas such
small droplets are not readily created with pipettes or pneumatic dispensers. In the case of contact printing, similar resolution to inkjet printing is achieved [12], but disadvantages

Fig. 1 Ink droplets (10 pL nominal droplet size) deposited at 100 m


drop spacing onto a glass slide using a piezoelectric inkjet printing device

Inkjet printed (bio)chemical sensing devices

Fig. 2 Ink droplets (10 pL nominal droplet size) ejected from an array
of nozzles with 21 m diameterdiameter; the distance between the blue
lines is 100 m

include the risks of spot-to-spot cross-contamination and of


damaging delicate substrates.
In (bio)chemical sensing device fabrication, screen printing
is a widespread method to produce electrodes on planar substrates [13]. The target material is deposited onto the substrate
by forcing the ink through the pores of the screen, which
defines the outline of the printed feature. For the fabrication of
printed electrodes or electrical contacts, screen printing is
probably the most serious competitor of inkjet printing,
since it enables very fast and low cost production with large
numbers of replicates. In some cases, screen printing is applied in combination with inkjet printing in order to profit
from the advantages of both technologies. From a material
deposition point of view, screen printing allows for the fast
creation of relatively thick films (20100 m). However, the
achievable control over the layer thickness is low compared
with inkjet printing. Furthermore, in terms of spatial control, a
separate screen is required for every print pattern, which
makes the technology useful for the reproduction of large
numbers of identical layouts, but lacks the flexibility offered
by inkjet printing when it comes to the free choice of shape
and layout for device prototyping.
Spin coating is a fast and simple method for depositing thin,
uniform layers of a material (e.g., a polymer film) on comparably large areas. However, it does not offer the possibility of
spatial control and, therefore, cannot compete with inkjet
printing in terms of resolution and flexibility. Furthermore, it
is not very economical in material use, since a significant
amount of the coating liquid does not remain on the substrate,
but is wasted.
Inkjet printing and the alternative material deposition
methods (1)(4) listed above are referred to as additive
deposition methods, although inkjet printing can also be
applied in a subtractive way [14, 15]. With the exception
of the solvent, only the amount of material required to obtain a
specific structural feature is deposited onto the substrate,
avoiding the waste of potentially high-cost materials. Their

5787

additive character makes those methods comparably economical in terms of material use. In contrast, photolithography [16]
is based on a subtractive approach, resulting in the creation of
waste. While photolithographic methods are superior to inkjet
printing in terms of achievable resolution, their subtractive
character calls for a larger number of processing steps and the
selective removal of pre-deposited material requires a specific
mask for every pattern, resulting in lower flexibility.
Material deposition methods featuring even higher resolution and patterning densities, such as for example soft lithography approaches [16, 17] and dip-pen nanolithography [16, 18],
are not discussed here, although they are increasingly applied,
in particular in the fabrication of bioanalytical devices.
However, their processing throughput and their cost efficiency
are at present not competitive to inkjet printing.

Piezoelectric or thermal inkjet printing? Consumer


desktop or research-use material printer?
Drop-on-demand (DOD) piezoelectric and thermal inkjet
printing are the most widely used technologies in industrial
process and consumer desktop printing, with the piezotechnology dominating the industrial area and the thermaltype the consumer office market [19]. Without providing
further technical details, the following are the basic principles
of these two methods: In piezoelectric printing, applying a
voltage to a piezoelectric element in contact with the printer
nozzle results in a volume displacement and the ejection of an
ink droplet. In thermal inkjet technology, the volume displacement is achieved by the creation of a gas bubble on a heating
element inside the nozzle. Judged from the number of application examples (Table 1), there still is a significant preference
for the piezoelectric technique when it comes to the fabrication of (bio)chemical sensors. In particular, when working
with protein containing printing inks, the fact of the ink being
exposed to a 300 C heat pulse seems to be detrimental to
enzyme activity. However, studies confirming the suitability
of thermal inkjet printing also for the printing of supposedly
sensitive enzyme inks have been published [2023]. On the
other hand, protein damage caused by shear forces occurring
in piezoelectric printer nozzles has been reported [24]. In
many situations, it is not necessarily the chosen printing
technology but rather the influence of ink additives (e.g.,
solvents, surfactants, viscosity modifiers, humectants), which
have a negative impact on the printing of protein containing
inks [25]. It is well known that enzyme denaturation occurs in
the presence of surfactants. Fortunately, the amount of surfactant required to reduce the surface tension of aqueous inks to
an ideally printable range (30 mN/m) is usually relatively low
and does not significantly decrease enzyme activity, in particular when using non-ionic surfactants (e.g., 0.1 wt% of Triton
X-100). Conversely, evaluated on the example of horseradish

Dimatix DMP-2811

Optical / colorimetric

Optical / absorbance

Electrochemical /
conductometric

Electrochemical /
chromoamperometric
Electrochemical /
resistance

Piezo

Epson R230

Organic solvent vapor,


temperature

NH3 gas

Acetic acid vapor

Ethanol, methanol vapor


NH3 gas
CO, H2, H2S, NO

Organic vapors
CO2

Alcohol vapor

H2O2, NADHe,
Salbutamol
Chemical O2 demand
(COD)
Ascorbic acid

Triglycerides

Epinephrine,
norepinephrine,
dopamine
Glucose, H2O2

Ethyl cellulose polymer sensing


layer with pH-indicator
Plasticized PMMAf sensing film with
chromogenic indicator
Monomer precursor for formation
of polymeric sensing film

Conducting polymer layer


(polyaniline)
Conducting polymer carbon
nanotube composite electrodes
Reduced graphene oxide electrodes
Pd-polymer nanocomposite as
seed-layer for electrode formation
PEDOT-PSS sensing layer
PEDOT-PSS sensing layer
Polyaniline nanoparticle sensing layer
WO3 nanoparticle sensing layer

Glucose oxidase or horseradish


peroxidase enzyme
Au/PEDOT-PSS nanocomposite
sensing layer
Graphene-PEDOT-PSS
nanocomposite sensing layer
TiO2 photoanode (nanoparticles)

Piezo

Dimatix DMP-2800

Au electrode array
(nanoparticles)
Au electrode precursor
(nanoparticles)
Polyaniline nanoparticle
sensing layer
Ammonium persulfate oxidant
for vapor phase polymerization
of polyaniline electrodes
PEDOT-PSSd film

Aqueous NH3

Piezo

Dimatix DMP-2800

Au electrodes (nanoparticles)

HP Deskjet D2360

Dimatix DMP-2831

Dimatix DMP-2811

HP Deskjet 693C
Epson C46/C48
Dimatix DMP-2800

HP 4250
Dimatix DMP-2800

HP Deskjet 690C

Thermal

Piezo

Piezo

Thermal
Piezo
Piezo

Thermal
Piezo

Thermal

Piezo

Piezo

Epson R290
Dimatix DMP-2800

Piezo

Dimatix DMP-2800

Piezo

Thermal

Olivetti

Dimatix DMP-2800

Thermal

Canon Pixma IP1300

Piezo

Piezo

anti C-reactive protein


(CRP) antibody
Interleukin-6
(cancer biomarker)
O2

Dimatix DMP-2800

Au and Ag electrodes
(nanoparticles)

Type

pH, glucose

Printerb

Electrochemical /
potentiometric,
amperometric
Electrochemical /
impedimetric
Electrochemical /
amperometric

Inkjet printed features

Sensing applications

Transduction mechanisma

Copy paper

Polymer optical waveguide

Polyester film
Silver electrodesc on PET film
Si/SiO2 substrates with
lithographically defined
Ti/Pt electrodes
LED lens surface

PET transparency
PET film

Screen-printed carbon
electrodes on filter paper
PET sheets

indium tin oxide glass slides

Carbon electrodesc

Carbon electrodesc

indium tin oxide glass


on cellulose acetate

[50]

Carbon electrodesc
on PET film
PET film

[62, 63]

[60]

[59]

[57]
[49]
[10]

[46]
[56]

[45]

[104]

[58]

[54, 55]

[53]

[20, 21]

[52]

[42]

[41]

[40]

[31]

Ref

Mixed cellulose ester

Polyimide

Coated paper

Coated paper

Substrate

Table 1 (Bio)chemical sensing devices fabricated by inkjet printing. Table items are grouped according to the sensor transduction mechanism, which is not necessarily the same order as in the main text

5788
N. Komuro et al.

pH

Optical / fluorescence

Mechanical / mass

Optical / colorimetric

Malathion, heroin,
cocain
pH, DNA, gas

Nitrite ions

pH, human IgG, mouse IgG

pH, protein, glucose

Streptavidin
NO2 gas

Rhodamine 6G

Optical / SERS

Carbohydrate-binding proteins

C-reactive protein (CRP)

Food spoilage
and ripening
Ricin

BSA, glucose
HRP activity

Malaria antigen
ATP, IgG

Hydrophobic barrier pattern


(reaction zones)
Detection antibody lines
Detection antibody lines (antibodies
and DNA aptamers coupled to
polymer nanoparticles)
Outline of microfluidic patterng
HRP enzyme inside
microfluidic channel
Sensing molecules
(oligodeoxyfluorosides)
Carbohydrate spots (galactose)
as sensing molecules
Photopolymerizable monomer
mixture including pH-indicator
Fluorophore labeled antibody
zone in microfluidic chip
BSAh-glycoconjugates
as sensing spots
Substrate hydrophobization
Raman enhancer layer
(Ag nanoparticles)
Raman enhancer layer
(Ag nanoparticles)
Alkanethiol SAMs, DNA
oligomers, polymer films
Protein layer (biotinylated BSA)
Microsphere-templated
BaCO3 films
Microfluidic pattern
Sensing spots with
chromogenic reagents, enzymes
Microfluidic pattern
Capture antibodies,
chromogenic reagents
Microfluidic pattern

Layer-by-layer assembled enzymatic


sensing film
Enzyme-doped sol-gel sensing
layer, chromogenic reagents,
oxidizing agent

Phenolic compounds

Neurotoxins,
organophosphate
pesticides, bacteria
in food samples
Phosphate

Inkjet printed features

Sensing applications

Optical / colorimetric
(fluorescence-based)

Transduction mechanisma

Table 1 (continued)

Canon Pixma ip4500

PicoJet-2000

PicoJet-2000

Dimatix DMP-2831
Hewlett-Packard

Autodrop MD-P-705-L

Epson Workforce 30

Thermal

Piezo

Piezo

Piezo
Thermal

Piezo

Piezo

Piezo

Piezo

Scienion S3 Flexarrayer
Epson Workforce 30

Piezo

Thermal

Piezo

Epson R280
Not-specified
research-type printer
Autodrop MD-P-705-L

Thermal

Thermal
Thermal

HP LaserJet 1000
Canon Pixma ip4500
HP Deskjet F4280

Piezo
Piezo

Thermal

Piezo

Piezo

Type

Scienion
Dimatix DMP-2800

Canon iP4700

Dimatix DMP-2800

Dimatix DMP-2800

Printerb

Filter paper

Filter paper

Filter paper

Silicon cantilevers
QCM

Silicon cantilevers

Chromatography paper

Chromatography paper

Microfluidic chip
(Si/ polydimethylsiloxane)
Silicon microring resonators

Optical fiber tip

Chromatography paper

Cotton paper

Filter paper
Filter paper

Nitrocellulose strips
Filter paper

Filter paper

Filter paper, cardboard paper

Filter paper

Substrate

[85, 86]

[84]

[15]

[76]
[107]

[75]

[74]

[73]

[72]

[71]

[70]

[69]

[68]

[106]
[23]

[93]
[105]

[89]

[6567]

[64]

Ref

Inkjet printed (bio)chemical sensing devices


5789

Glucose

pH
IgG

Electrochemical / resistance
Optical / reflectometry

Photonic crystals
Antibody spots
Blocking solution

Microfluidic structure

Hydrophobic barrier pattern


(reaction zones)c

Sensing spots with


chromogenic reagents
Microfluidic pattern
Sensing spots with pH-indicators
Enzyme-doped sol-gel sensing
layer, chromogenic reagents
Hydrophobic barrier
Microfluidic pattern

Inkjet printed features

Italic font indicates the application of a research-use material printer, whereas plain font indicates the use of a consumer desktop printer.

2-(Dibutylamino)ethanol.

Bovine serum albumin.

The actual pattern is drawn with a wax-pen, following the inkjet printed pattern outline.

Poly(methyl methacrylate).

Nicotinamide adenine dinucleotide.

Poly(3,4-ethylenedioxythiophene/polystyrene sulfonic acid).

Electrodes are fabricated by screen-printing.

Piezo

Piezo

Dimatix DMP-2800

PicoSpot Jet
Dispensing System

Thermal

Piezo

Piezo

Type

Canon Pixma ip4500

Dimatix DMP-2800

Epson PX-101

Printerb

Bold face indicates an all inkjet printed approach, where all major sensor fabrication steps were achieved with an inkjet printer.

DBAEi, NADHe

Optical / electrochemiluminescence
Optical / colorimetric

Heavy metals

pH, H2O2

Sensing applications

Optical / colorimetric

Transduction mechanisma

Table 1 (continued)

Hydrophobically
coated glass slides

Various paper substrates

Filter paper

Filter paper

Filter paper

Substrate

[108]

[91]

[90]

[92]

[87, 88]

Ref

5790
N. Komuro et al.

Inkjet printed (bio)chemical sensing devices

peroxidase (HRP), it was found that the presence of viscosity


modifiers can lead to a significant decrease in enzyme activity,
in particular when using high molecular weight additives. In
the case of poly(ethylene glycol) addition, an increasingly
negative impact on HRP activity with increasing molecular
weight of the poly(ethylene glycol) was observed. The use of
the charged polymer sodium carboxymethyl cellulose was
shown to have the lowest impact on enzyme activity in the
case of HRP. This is due to the high viscosity modifying
power of that material, making the application of small
amounts sufficient (e.g., 0.5 wt% of carboxymethyl cellulose
are sufficient to achieve an ink viscosity of 5 cPs, compared
with up to 80 wt% required to obtain the same viscosity in the
case of using ethylene glycol as additive).
As a consequence, the type of inkjet printing technology
applied is not of primary concern. Both piezoelectric and
thermal inkjet printing can be considered for the prototyping
and mass production of (bio)chemical sensing devices. With
both methods, the evaluation of the stability of a particular
protein/enzyme under inkjet printing conditions is required.
Probably researchers will tend to stick with piezoelectric
systems, as long as those retain their dominance in the
industrial inkjet printing sector.
The question of what type of printer to use can be
addressed from several points of view. Factors to consider
include cost, user-friendliness, and versatility. While the
more sophisticated material printing systems targeting research applications offer full control over printing parameters in terms of hardware (e.g., choice of nozzle diameter,
heatable nozzles and printing stages, visual confirmation of
ink droplet formation, etc.) and software (e.g., control of
piezo-pulse voltage and duration, etc.), consumer desktop
printers are basically black-box systems offering only
limited user control via the original printer driver and
graphics software used for printing pattern creation (e.g.
print quality selection, etc.). On the other hand, the simplicity of these devices allows for the adaptation to specific
needs, which is less possible with the high-cost material
printers. In addition, all current consumer desktop printers
come equipped with at least four ink tanks (black, cyan,
magenta, yellow), thus theoretically allowing the simultaneous use of multiple inks. Although this feature is rarely
used in the printing of (bio)chemical sensing devices with
consumer desktop printers up to now, it is a potentially large
advantage over many material printing systems, most of
which can handle only a single type of ink at a time. A
strong point in favor of research-use material printers is their
resistance to a variety of chemical substances, including
organic solvents of high dissolving power. Among the material printing instruments available on the market, those
offering single-use exchangeable print heads and cartridges might be advantageous when working with experimental printing inks that tend to clog printer nozzles,

5791

especially by particle aggregation. Users of such difficult


inks might feel more comfortable working with such a
system, compared with a high precision print head instrument costing several thousand dollars.
Many research groups rely on several types of inkjet printing systems (Fig. 3). In our own work, a desktop printer is
given preference whenever applicable. One argument besides
cost for favoring the use of consumer inkjet printers is the idea
that if a (bio)chemical sensing device can be fabricated on an
unmodified consumer desktop printer, it is probably adaptable
for mass production on any type of currently applied inkjet
printing system. The consumer inkjet printer quasi provides
the proof-of-concept for the general printability of a sensing
device.

(Bio)chemical sensing devices fabricated by inkjet


printing
In Table 1, a nonexhaustive list of application examples of
inkjet printing technology (both desktop and material printers)
from the scientific literature dealing with (bio)chemical sensing devices is presented. Not all entries listed in the table are
discussed in detail. The table as well as the ensuing text
sections have been organized based on the respective sensing
transduction mechanism, with overlaps occurring. As a result,
it is readily visible that inkjet printing technology is nowadays
applied to the development and fabrication of almost any type
of (bio)chemical sensing device.
Electrochemical sensors
Promoted by the progress in printed electronics in general,
DOD inkjet printing technology is widely used to fabricate
electrodes, resistors, transistors and other electronic parts for
electrochemical sensing devices. The most regularly explicitly
stated reasons for referring to inkjet printing in these applications are the low cost, the simple process, the efficient use of
ink and the compatibility with various substrates.
Electrode fabrication
Because of its simplicity and suitability for mass production,
screen printing has become a major technology for the fabrication of planar electrodes, which form the basis for a variety
of electrochemical sensors [13]. Widely applied screen printed
electrode materials are carbon, silver and gold. However, besides of lacking flexibility by requiring a specific screen for
every electrode pattern, the loss of ink in particular in smallscale fabrication is a further disadvantage of this method.
Therefore, DOD inkjet printing technology with its efficient
use of ink has become an attractive alternative for the fabrication of electrodes, especially for research and prototyping

5792

N. Komuro et al.

Fig. 3 Different types of inkjet printing devices for laboratory use: (a) research-use material printer accommodating a multiple nozzle single-use
print head; (b) research-use inkjet dispenser with single nozzle print head; (c) piezoelectric consumer desktop printer

applications, where pattern flexibility is highly desired.


Because metal electrode fabrication is the application of inkjet
printing technology closest to routine in (bio)chemical sensor
development, some work reported in this area is discussed in
more detail.
In particular, commercially available silver nanoparticle
suspensions (silver nanopaste) are already routinely used
for Ag electrode printing in the fabrication of (bio)chemical
sensors, for example [2633]. The applied silver nanopaste
inks are based on organic solvents and dispersing agents to
prevent aggregation of the metal nanoparticles. Their printing requires the use of a research-type material printer. In
addition, to gain sufficient conductivity of the printed metal
traces, an elevated temperature sintering process (>200 C)
is necessary after printing to remove the solvent and the
particle stabilizers, which prevents the use of heat sensitive
substrates. These limitations are the driving force for continued research on alternative methods.
In a very simple, environmentally friendly approach,
Bidoki et al. fabricated silver electrodes on paper, overhead
projector transparencies, and cotton, by subsequently printing
aqueous solutions of ascorbic acid and silver nitrate with a
thermal desktop printer [34]. When the silver nitrate and the
ascorbic acid mix on the substrate, silver nitrate is reduced to
metallic silver. Unreacted materials are removed by washing
with water, followed by heating (150 C, 20 s). An additional
advantage of this approach, besides of being solvent-free,
requiring no extended sintering, and being printable on a
desktop printer, is that with the absence of suspended particles
in the ink, the risk of printer nozzle clogging is much less an
issue. A serious drawback, however, is the low conductivity,
which is only about 0.3% the value of bulk silver. For this
reason, researchers keep searching for methods to reduce the
temperature and time required for sintering of structures
printed from silver nanoparticle inks, or to increase the conductivity of silver traces obtained by the printing of metal salt
precursors. An interesting example of the first approach is the
work by Magdassi et al., where the sintering temperature for
an inkjet deposited silver nanoparticle ink has been lowered
down to room temperature [35]. This has been achieved by
printing Ag nanoparticles (thermal desktop printer) stabilized
by a negatively charged polymer, poly(acrylic acid) sodium

salt, onto paper or plastic substrates precoated with a positively charged polymer, poly(diallyldimethyl ammonium chloride). The induced charge neutralization results in a
spontaneous sintering of the silver nanoparticles at room
temperature, leading to metallic silver traces with 20% the
conductivity of bulk silver, a value otherwise only achieved
by extended sintering at high temperatures. Perelaer et al. have
demonstrated a time reduced sintering method for electrodes
printed with a commercial silver nanopaste ink (piezoelectric
research-type printer) resulting in 40% of bulk silver conductivity [36]. The method relies on a combination of photonicand microwave-induced sintering that can be completed in
less than 15 s, making it suitable for a continuous roll-to-roll
fabrication process. In another approach chosen by Tobjrk et
al., conductivities corresponding to 10%20 % of that of bulk
silver have been obtained by exposing electrodes printed with
a commercial Ag nanopaste ink (piezoelectric research-type
printer) for 15 s to an IR lamp [37]. This method was also
successfully applied to electrodes printed from gold nanoparticle inks (piezoelectric research-type printer) [38]. The same
research group has finally combined inkjet printed gold working and counter electrodes with an inkjet printed silver electrode, electrochemically converted into an Ag/AgCl reference
electrode, to obtain an amperometric three-electrode paper
sensing device for glucose [31]. This low-cost inkjet printed
three-electrode paper chip platform, which can be obtained
through a roll-to-roll process, performs equally well to a more
costly conventional electrochemical sensor setup. Nie et al.
have demonstrated the possibility to achieve highly conductive silver lines on PET substrates (50% the conductivity of
bulk silver) from a particle free metal precursor solution
(silver citrate) [39]. The approach is based on the thermal
reduction of the inkjet deposited silver citrate (piezoelectric
desktop printer) in the presence of 1,2-diaminopropane. The
formation of a silver-amine complex results in a decreased
redox potential, allowing the thermal reduction of Ag+ at
relatively low temperatures (50 min at 230 C). By reducing
the heat treatment temperature to 150 C, which allows the use
of more heat-sensitive substrates, still 10% of bulk silver
conductivity can be achieved.
In many applications, the use of gold electrodes is preferable over silver electrodes because of the chemical inertness of

Inkjet printed (bio)chemical sensing devices

this metal. Commercially available or laboratory prepared Au


nanopaste inks are generally composed of a high boiling
organic solvent and alkanethiol capped Au nanoparticles
[38]. Therefore, they suffer from limitations similar to Ag
nanopaste inks. However, gold electrodes allow for very simple surface modification with thiol compounds, making them
particularly useful for the immobilization of capture molecules, which has, for example, been applied in the development of inkjet printed electrochemical immunosensors
(piezoelectric research-type printer) [40, 41]. The inks for
the printing of these electrodes contain 15 wt% of gold
nanoparticles. The use of an organic solvent and of the capping reagents is necessary to allow for the stable dispersion of
the metal nanoparticles at such high concentrations. Hu et al.
have realized an interesting solution to overcome this drawback. For the fabrication of their paper-based solid-state electrochemical oxygen sensor, they printed a diluted aqueous Au
nanoparticle ink from a piezoelectric desktop printer onto a
mixed cellulose ester membrane [42]. The amount of gold
nanoparticles deposited by the 7-fold printing of the dilute
aqueous ink is not sufficient to achieve electrically conducting
metal traces. However, the printed features act as nucleation
sites for the following seeded growth of metallic gold upon
immersion of the substrate into a HAuCl4 containing plating
solution. After seven print cycles, followed by eight growth
cycles, a conductivity of approximately 10% of that of bulk
gold was achieved. With this method, the authors were able to
produce more than 200 electrode arrays in 30 min of printing,
followed by 2 h of Au growth with no other equipment than a
common desktop inkjet printer (Fig. 4).
While the fabrication of metallic Ag electrodes (and to a
lesser degree also of Au and Cu electrodes) by inkjet printing is becoming a routine procedure, the situation is different for another important electrode material. For carbon
electrodes that very often form the basis of electrochemical
(bio)sensors, screen printing continues to remain the fabrication technology of choice. The probably most impressive
success story of screen printed carbon electrodes is their
widespread application as working electrodes in disposable
test strips of personal glucose monitoring devices for patients suffering from diabetes [13, 43].
A few inkjet approaches with carbon nanotube-based
inks have been reported, since the printability of this material from aqueous dispersions has been first demonstrated
[44]. One example is an inkjet printed (thermal desktop
printer) alcohol vapor sensor relying on conductivity
changes of carbon nanotube electrodes, stabilized with a
conducting polymer, upon exposure to alcohols [45]. The
use of a different carbon material, an aqueous dispersion of
reduced graphene oxide, has first been demonstrated for the
fabrication of an inkjet printed (thermal desktop printer)
electrochemical gas sensor on standard overhead projector
films [46]. Exposure to organic vapors results in changes of

5793

Fig. 4 Inkjet printing of gold electrode arrays on mixed cellulose ester


membranes for fabricating paper-based electrochemical O2 sensors: (1)
inkjet printing of Au nanoparticle electrode precursor patterns, (2)
growth of Au nanoparticle patterns into Au electrode arrays, (3) cutting
a paper-based Au electrode array (PGEA) from its ensembles, (4)
electric connection and size control of a PGEA, and (5) addition of
BMIMPF6 ionic liquid electrolyte from the back side of a PGEA to
fabricate the final O2 sensor. Reprinted with permission from Reference [42] Hu C, Bai X, Wang Y, Jin W, Zhang X, Hu S (2012) Inkjet
printing of nanoporous gold electrode arrays on cellulose membranes
for high-sensitive paper-like electrochemical oxygen sensors using
ionic liquid electrolytes. Anal Chem 84:37453750. Copyright
(2012) American Chemical Society

the electrode conductivity. These are examples of applications where the inkjet deposited electrode material simultaneously acts as the active sensing layer.
Inkjet printing of active sensing layers for electrochemical
sensors
It has been shown in the previous section that inkjet printing
can be a great tool for the very simple and rapid fabrication
of electrodes, in some cases requiring nothing more than an
off-the-shelf consumer desktop printer. Another important
application of inkjet printing technology is the coating of
electrodes with an active sensing layer. Frequently, that
layer consists of a film of a conducting polymer with the
purpose of fabricating an electrochemical gas sensor
whereby, in some cases, the polymer simultaneously
acts as electrode and active sensing layer without the
requirement of a separate underlying electrode substrate.
Weng et al. have reviewed the printing of conducting polymers in general [47].
Widely inkjet printed conductive polymers include
polypyrrole, polyaniline (PANI), and poly(3,4-ethylene
dioxythiophene)-poly(styrenesulfonate) (PEDOT-PSS).
One problem encountered with PANI is the fact that it is
insoluble in water and even in many common organic solvents, which makes the straightforward formulation of
inkjet printable inks challenging. When the use of an offthe-shelf desktop printer is targeted, the inks have to be

5794

based on water or mild organic solvents. One way to resolve


this defect is to use dodecylbenzenesulfonic acid (DBSA)stabilized PANI nanoparticles for inkjet printing. The anionic dopant DBSA prevents the aggregation of PANI, which is
the main cause of printer nozzle clogging. It allows for the
preparation of aqueous dispersions with a relatively high
polymer content [48]. DBSA-PANI dispersions stabilized
against aggregation can be printed not only from state-ofthe-art research printers but also from desktop printers.
Crowley et al. reported sensors for aqueous or gas phase
ammonia detection obtained by a combination of screen
printing of underlying silver or carbon electrodes and inkjet
printing of DBSA-PANI on a piezoelectric research-type or
a piezoelectric desktop printer [49, 50]. The thickness of the
deposited PANI chemiresistor films significantly influences
the gas sensor properties (e.g., response time, stability, reproducibility, etc.). A great advantage of the inkjet printing
approach is the flexible variation in the deposited total film
thickness by simply varying the number of applied printing
layers, which under the conditions used by the authors had a
thickness of 170 nm [51].
A different approach to a PANI pattern printed from a
thermal desktop inkjet printer has been reported for a biosensor targeting the detection of electrochemically active biomolecules excreted from cells [52]. In this example, only an
aqueous solution of an oxidant, ammonium persulfate, is
inkjet patterned on the substrate. Doped PANI films are then
created through vapor phase polymerization by exposing the
patterned substrate to aniline monomer/HCl at 70 C for 15
min. With this method, dispersion stability of the printing ink
is not an issue, since a simple aqueous solution is the only
inkjet deposited material (Fig. 5).
Having advantages such as high conductivity and high
optical transparency, PEDOT-PSS is widely used in electrochemical and optical sensing. The polymer is water-soluble
and ready-to-print aqueous ink dispersions are commercially
available. The research group of Setti and coworkers was
among the first to use this conducting polymer for the fabrication of inkjet printed amperometric enzymatic biosensors on
a modified thermal desktop inkjet printer [20, 21]. They inkjet
printed glucose oxidase and horseradish peroxidase solutions
on conducting polymer layers obtained by inkjet deposition of
aqueous PEDOT-PSS dispersions on indium tin oxide glass.
In an amperometric enzymatic biosensor for triglyceride detection, Phongphut et al. used a novel composite material of
Au nanoparticles and PEDOT-PSS inkjet printed on screen
printed carbon electrodes (SPCE), where the Au nanoparticles
enhanced the electrochemical response [53]. Another composite material enhancing the electrochemical activity of a
PEDOT-PSS amperometric sensor has been presented by
Tuantranont and coworkers for the direct electrochemical
sensing of hydrogen peroxide (H2O2), nicotinamide adenine
dinucleotide (NADH), and salbutamol, respectively [54, 55].

N. Komuro et al.

An electrochemically synthesized graphene-PEDOT-PSS


nanocomposite was inkjet printed as an aqueous dispersion
on top of screen printed carbon electrodes. Sensors prepared
with the composite material showed significantly higher electrochemical response compared to unmodified SPCE or
PEDOT-PSS modified electrodes.
Tseng et al. made use of a piezoelectric research-type
inkjet printer for both the patterning of Au electrodes and
the coating with a PEDOT-PSS film for the development of
a CO2 gas sensor (Fig. 6) [56]. This work is one example for
using inkjet technology for both the electrode fabrication
and deposition of the active sensing layer.
Mabrook et al. fabricated a chemiresistor for alcohol sensing by printing PEDOT-PSS on polyester films [57]. They
noticed different response behavior depending on the
conducting polymer film thickness, which was easily controlled by the number of printing layers. For thick films (4-5
print layers, 500-700 nm thickness), reversible response to
ethanol vapor was found. On the other hand, thin films (single
print layer, about 100 nm thickness) behaved as chemical
fuses, responding irreversibly above a certain alcohol vapor
concentration threshold. These differences can be explained
by the solubility of PEDOT-PSS in alcohols, resulting in
structural rearrangements of the sensing films upon exposure
to the analyte. Compared to a sensing layer obtained by drop
casting, a thin inkjet printed film has a more disordered and
opened structure, since it is formed by single partially
overlapping droplets. The rapid penetration of alcohol vapor
into an open structure thin film results in the irreversible
breakdown of electrical conductivity. This example demonstrates that different sensor response can be achieved from the
deposition of a single polymer film material at different densities and thicknesses.
The formation of controlled particle layers of semiconducting metal oxides is of special interest for the development of
electrochemical gas sensors. Kukkola et al. printed additivefree aqueous suspensions (piezoelectric research-type printer)
of catalyst-modified WO3 nanoparticles for CO, H2, H2S and
NO sensing [10]. Owing to the extremely efficient material use
of the inkjet printing technology with almost no waste of ink
compared to other material deposition methods, the authors
calculated the material costs (only the WO3 nanomaterial) for
their devices to be no more than 1 Euro per 1 million devices.
They concluded that by combining inkjet deposition with low
cost substrates, the costs of a final sensor would be mostly
determined by the measurement electronics, rather than the
sensing materials.
Yang et al. have selected inkjet printing as a cost-effective
mass production technology to produce robust and highly
reproducible TiO2 photoanodes, applied to the measurement
of glucose, phenol, potassium hydrogen phthalate, glutaric
acid, malonic acid, and chemical oxygen demand [58]. They
were particularly looking at a replacement for the dip coating

Inkjet printed (bio)chemical sensing devices

5795

Fig. 5 Schematic diagram of an inkjet printing-assisted fabrication of


a polyaniline (PANI) pattern. Reprinted from Reference [52] Oh W-K,
Kim S, Hwan Shin K, Jang Y, Choi M, Jang J (2013) Inkjet printed

polyaniline patterns for exocytosed molecule detection from live cells.


Talanta 105:333339, Copyright (2013), with permission from
Elsevier

technique, which lacked in reproducibility. They used a piezoelectric consumer inkjet printer to deposit an aqueous colloidal TiO2 particle (6 wt%) ink containing 1.8 wt% of
carbowax as viscosity modifier onto indium tin oxide glass.
The inkjet printed sensing layers were directly compared with
those deposited by the conventional dip coating method.
Figure 7 shows a comparison of inkjet printed and dip coated
TiO2 electrodes. A more homogenous film formation is clearly confirmed for the inkjet printed electrode. As in other cases
of depositing an active sensing layer by inkjet printing, film
thickness control was easily achieved by varying the number
of printed layers. The sensing results revealed that inkjet
printed TiO2 films have good fabrication reproducibility and
that the relative standard deviations for independently fabricated electrodes are smaller compared with commonly used
dip coated sensors.

cost effective, inkjet printing technology is increasingly being


used as a deposition method for active layers of optical
(bio)chemical sensors.
Absorbance/colorimetry-based sensors

Research on optical (bio)chemical sensors has been very intensive over the past decades and continues to increase. The
essential part of any optical sensor is the active sensing layer
incorporating the signal transducing molecules. Major issues
with mass production of optical sensors are reproducible,
speedy, and controlled deposition of the active sensing layer.
Regarded as being rapid, reproducible, controllable (e.g., size,
shape, position of deposited material), contamination-free, and

Absorption spectrometry and colorimetry are among the


simplest optical detection methods. Inexpensive optical
components such as light emitting diodes (LEDs) are nowadays widely available and applied to field-deployable optical (bio)chemical sensing devices. To fabricate sensitive
and reproducible devices, the uniformity of the sensing layer
deposited onto an LED is one of the most important factors.
All of the deposition methods discussed earlier, including
dip coating, drop casting, spin coating, and screen printing,
with their known deficiencies in terms of reproducibility,
resolution, freedom of patterning and others, are regularly
reported for the fabrication of optical sensors. Inkjet technology is regarded as a painless solution to overcome these
deficiencies, and resulting sensors are assumed to have
greatly improved reproducibility and sensitivity. In the following, a few selected examples will be discussed to verify
this assumption.
M. OToole et al. reported a sensor for gaseous acetic
acid based on an ethyl cellulose sensing layer incorporating
a colorimetric pH-indicator [59]. They dissolved all sensing
materials in an organic solvent (1-butanol) and inkjet printed

Fig. 6 Schematic illustration of the fabrication process of a PEDOT


PSS film modified electrode, where both the electrode pattern and the
conducting polymer film are deposited by inkjet printing. Reprinted
from Reference [56] Tseng CC, Chou YH, Hsieh TW, Wang MW, Shu

YY, Ger MD (2012) Interdigitated electrode fabricated by integration


of ink-jet printing with electroless plating and its application in gas
sensor. Colloid Surf A, 402:4552, Copyright (2012), with permission
from Elsevier.

Optical sensors

5796

Fig. 7 Comparison of an inkjet printed TiO2 electrode (left) and dipcoated electrode (right). Reprinted from Reference [58] Yang M; Li L.H.;
Zhang S.Q.; Li G.Y.; Zhao H.J. (2010) Preparation, characterization, and sensing application of inkjet printed nanostructured TiO2
photoanode. Sens Actuators B 147:622628, Copyright (2010), with
permission from Elsevier

(piezoelectric research-type printer) the solution directly


onto the lens surface of an LED light source. For comparison, sensors were also prepared by drop casting 5 L of the
printing ink onto the LED by means of a micropipette. Inkjet
deposition allowed for the reproducible control of sensing
film thickness by varying the number of printed 100 nm
thick layers. The direct comparison of 10 repetitions of
inkjet printed and drop-casted sensing films (Fig. 8) clearly
demonstrates the superiority of inkjet printing in terms of
sensor fabrication reproducibility. The relative standard deviations in the response signal of 10 independently deposited sensing layers were 5.6% and 68.0% for the inkjet (seven
print layers) and drop casting approach, respectively. These
results demonstrate an excellent improvement of fabrication
reproducibility, especially when considering that seven
layers had to be inkjet printed in order to achieve a sensor
signal intensity comparable to the drop casted devices.
Courbat et al. have applied inkjet printing (piezoelectric
research-type printer) of an organic solution of sensing
components (pH-indicator, polymer matrix, plasticizer) onto
a low-cost polymer waveguide to develop an optical ammonia gas sensor [60]. The thickness of the printed film was
140 nm. The limit of detection (LOD) of this sensor was
calculated as 104 ppb. The same group has earlier reported
an ammonia gas sensor fabricated by spin coating an organic solution of identical sensing materials onto a planar glass
waveguide [61]. The LOD of the spin-coated sensor with
159 nm film thickness was found to be 2 ppb. The authors
have demonstrated that inkjet printing is a reasonable alternative to spin coating for the deposition of a polymeric
sensing layer onto a planar waveguide. Although a deterioration of the LOD has been observed, inkjet technology is
more suitable for mass-production by roll-to-roll processes

N. Komuro et al.

compared with spin coating. But this example also shows


that inkjet printing applied for the deposition of a sensing
film does not always result in improvements of sensor
performance.
In a significant number of optical sensors, the active sensing
components are immobilized into an organic polymer matrix.
Many of the applied polymer matrices are insoluble in water.
Therefore, research-type printers able to withstand organic
solvents have to be used for their inkjet-based deposition. In
addition, even when using a solvent-resistant printing system,
intrinsic limitations of inkjet printable compositions regarding
viscosity, density, and surface tension do apply. As a consequence, only inks with relatively low polymer concentrations
can be used, requiring repeated printing cycles to deposit a
sufficient amount of active sensing material. A large number of
printing cycles is time-consuming and results in lower sensor
fabrication reproducibility. As in the case of active sensing
layers for electrochemical sensors, there is a desire to deposit
polymeric membranes from relatively concentrated printing
inks based on aqueous media.
Yoon et al. demonstrated one example of overcoming this
limitation [62]. Polydiacetylenes have interesting properties as
optical sensing materials, because of their stimulus responsive
color and fluorescence changes promoted by heat, mechanical
stress, environmental, chemical and biological interactions.
Unfortunately, neither polydiacetylenes, nor their monomeric
precursors are soluble in water. But the low cost commercially
available 10,12-pentacosadiynoic acid monomer and its derivatives are known to form self-assembled vesicular structures in aqueous environment, and were selected as inkjet
printable precursors. However, the vesicles tend to aggregate,
which limits the concentration of the printing inks to about 5
mM. The authors looked at the application of commercially
available ionic and nonionic surfactants to stabilize the vesicles by co-assembly with the monomer molecules. With simple 10,12-pentacosadiynoic acid and the nonionic surfactant
Brij78, long-time stable aqueous printing ink suspensions
with up to 13 mM of monomer were obtained and printed
on the surface of unmodified copy paper by a thermal desktop
printer. Conversion into a polymeric structure was achieved
by UV irradiation with a handheld UV lamp during 3 min.
Chemical sensors with colorimetric response to organic solvents (THF, hexane) and pH were demonstrated, among
others. In a later report, the same research group has extended
the system to develop microemulsion-based inks containing a
diacetylene monomer, surfactant, co-surfactant, and organic
solvent [63]. This approach allowed the stabilization of organic droplets containing a high concentration of diacetylene
monomer in aqueous solutions. The application to a paperbased colorimetric temperature sensor was demonstrated.
The conservation of enzyme activity is an important issue
in the development of enzyme-based biosensors. The possible influence of the selected printing technology (thermal or

Inkjet printed (bio)chemical sensing devices

5797

Fig. 8 Demonstration of superior fabrication reproducibility of inkjet


deposited sensing layers: plot obtained from multiple (n = 4) injections
of gaseous acetic acid employing 10 (a) drop casted and (b) inkjet
printed colorimetric sensors. Reprinted from Reference [59] O'Toole

M, Shepherd R, Wallace GG, Diamond D. (2009) Inkjet printed LED


based pH chemical sensor for gas sensing. Anal Chim Acta 652:308
314, Copyright (2009), with permission from Elsevier

piezoelectric) has already been discussed above. In the


following, two examples of enzyme-based inkjet printed
biosensor approaches with colorimetric signal detection are
presented. In all cases, the enzymes are chemically
immobilized onto the surface of paper substrates, and the
same piezoelectric research-type inkjet printer has been
applied.
In the first example, Alkasir et al. reported a biosensor for
the colorimetric detection of phenolic compounds [64]. The
sensing layer was build up in a layer-by-layer (LbL) format
by sequential inkjet printing of aqueous chitosan, sodium
alginate, and tyrosinase enzyme solutions onto a filter paper
substrate. The sensor response relies on the color change
caused by the specific binding of quinone formed by enzymatic conversion of the phenolic analyte to the immobilized
chitosan. The authors managed to demonstrate that inkjet
printing is a useful tool for the large-scale production of
LbL-based sensors on paper substrates, while retaining enzyme activity and functionality. Although it was experimentally observed that the sensitivity (color intensity change per
sample concentration unit) of the inkjet printed sensors was
slightly lower than that of paper disks LbL coated by manual pipetting, the sensor performance can be regarded as
comparable. The authors of this study have not optimized
the number of inkjet printed layers for each deposition step,
while an optimization was performed for the manual
pipetting approach. Therefore, it cannot be stated that the
observed reduction in sensitivity is directly related to the
inkjet printing process (e.g., enzyme denaturation).
Hossain et al. have reported sol-gel-based colorimetric
sensors on paper substrates for the detection of neurotoxins
and pesticides [65, 66]. Both systems rely on the inhibition
of acetylcholine esterase in the presence of the analyte,
which is converted into a colorimetric signal by using chromogenic enzyme substrates. The authors put particular emphasis on the permanent and stable immobilization of the
enzyme on the paper surface [65]. In their work, the applied
enzyme is immobilized on the paper surface by sandwiching

between two layers of a sodium-silicate sol-gel material


deposited on top of a poly(vinylamine) underlayer. The
three materials are separately and sequentially inkjet printed
on the paper substrate. No loss in enzyme activity was
observed after storage of printed sensors at 4 C for 1
month. The authors have optimized all three printing inks
(sol-gel ink, enzyme ink and poly(vinylamine) ink) for
inkjet printability by varying the amount of glycerol added
as viscosity control agent. Excellent printability was
achieved upon the addition of 30 wt% of the viscosity
modifier to the enzyme ink. More than 95% of the enzymatic activity of acetylcholine esterase was maintained with such
ink composition. This stands in contrast to an observed loss of
enzyme activity of 30%, reported by Di Risio et al. for an
HRP-based enzyme ink with 30 wt% of glycerol as viscosity
modifier [25]. This significant difference in impact of an
identical viscosity modifier on enzyme activity is another
indication for the requirement to evaluate the stability of each
particular enzyme under specific inkjet printing conditions, as
has already been mentioned further above.
The same sodium-silicate sol-gel material ink has also
proven useful for the immobilization of other sensing reagents, such as chromogenic enzyme substrates or an oxidizing agent, as has been demonstrated in paper-based inkjet
printed lateral-flow devices for the detection of bacterial
contamination of food samples [67]. All materials required
for the sensor have been deposited on filter paper by piezoelectric inkjet printing of aqueous media. Specific enzymes
present in lysed bacteria result in the development of a
colored line in the case of contaminated samples.
Fluorescence-based sensors
For (bio)chemical sensors inkjet printed on paper substrates,
colorimetry is preferred due to the extremely simple signal
detection (e.g., by the naked eye). However, there are reports
of fluorescence-based systems where, for example, a fluorescence microscope is used as the signal detector. Kwon et al.

5798

have realized the detection of food spoilage or the monitoring


of fruit and vegetable ripening by headspace analysis with a
sensor array consisting of inkjet printed fluorescent labeled
DNAs on paper [68]. DNA-like oligodeoxyfluorosides
(ODFs), where DNA bases are replaced by fluorophores, have
been deposited on a paper substrate as aqueous solutions by a
thermal desktop printer. The authors obtained uniform sensing
spots of reproducible brightness by optimizing their printing
inks in terms of viscosity to control the spreading behavior on
the substrate. They have also investigated a number of different substrates and concluded that brightener-free cotton paper
was the most suitable material. Only 25 pmol of ODF was
required per sensing test. The deposited ODFs directly
responded by fluorescence signal increases or color shifts
when exposed to volatiles released from the samples, allowing
for colorimetric data processing of fluorescence micrographs.
As in other examples, the authors explain their selection of an
inkjet printing approach in particular with its convenience,
simplicity, effectiveness, and low costs.
In another inkjet-based paper sensor relying on colorimetric or fluorescence transduction, Yu et al. have developed a
method for the direct covalent immobilization of small molecules, proteins, and DNA used as sensing materials for biosensors [69]. In one application example, galactose is
deposited for covalent attachment on a divinyl sulfone
pretreated filter paper by a piezoelectric desktop printer, and
visualized with fluorescently labeled ricin.
In the following, three examples of fluorescence-based
systems are presented, where the spatial resolution and the
droplet formation reproducibility of the inkjet printing technology are essential factors. However, it is needless to say
that in terms of manufacturing, there is no basic difference
between absorbance/colorimetry based (bio)chemical sensors and those relying on fluorescence or other signal transduction, when it comes to the need for precisely depositing a
functional material on a sensor platform.
Already in 2006, Carter et al. have used a piezoelectric
research-type inkjet setup for the deposition of a fluorescent
pH-indicator dissolved in a photopolymerizable monomer
mixture in the form of multiple dots on the tip of a 500 m
diameter optical fiber [70]. The authors relied on the excellent
spatial control and reproducibility of the method. The obtained
dots showed very small variations in diameter (92.2 2.2 m),
height (35.0 1.0 m) and roundness. Up to seven dots were
independently placed on a single fiber tip, without having
contact with each other (Fig. 9).
Gervais et al. used a piezoelectric research-type printer to
place detection antibodies required in an fluorescence-based
immunodiagnostic assay for C-reactive protein (CRP) onto
a microfluidic capillary device [71]. As in the previous
example, precision is the most essential reason for selecting
an inkjet printing method. A total of 3.6 nL of detection
antibody solution in the form of 20 drops of 180 pL was

N. Komuro et al.

deposited into an area of the device that is only 100 m


wide.
Kirk et al. applied a piezoelectric research-type printer for
the multiplexed deposition of bioactive reagents onto silicon
micro-ring resonator devices intended for label-free protein
detection [72]. The authors referred to inkjet printing to replace the conventionally used fluidic masking techniques or
pin-contact printing methods because they suffer from limited
resolution (low functional density), requirement of large reagent volumes, or damage caused by contacting the waveguide surface. The inkjet printer deposition accuracy was
calibrated by printing fluorescently labeled proteins, followed
by fluorescence microscopic analysis of deposited spots. By
modifying the printers software setting, a final spot precision
of 5 m was achieved. Besides ink and printer parameter
optimization towards reproducible droplet formation, the reproducible fixation of the substrate on the stage of the printer
is an additional requirement when such high spatial resolution
is targeted. The authors used a precisely machined chip holder
for their experiments. In an impressive performance of rapid
and efficient mass-production of densely functional microring resonator arrays by inkjet printing, 10 sensor chips were
printed, each containing six evenly spaced micro-rings (30
m diameter), in a total of 9 s consuming less than 25 nL of
reagent. Finally, precise coating of individual micro-rings with
different protein conjugates without any cross-reactivity was
demonstrated. With their results, the authors have proven that
inkjet printing can overcome the current limitations of silicon
micro-ring resonator devices in terms of fabrication throughput and functional density.

Fig. 9 Individual polymeric microdots (92 m diameter) incorporating


fluorescein as pH indicator deposited on the tip of a 500 m diameter
optical fiber bundle. Reprinted from Reference [70] Carter JC, Alvis RM,
Brown SB, Langry KC, Wilson TS, McBride MT, Myrick ML, Cox WR,
Grove ME, Colston BW. (2006) Fabricating optical fiber imaging sensors
using inkjet printing technology: a pH sensor proof-of-concept. Biosens
Bioelectron 21:13591364, Copyright (2006), with permission from
Elsevier

Inkjet printed (bio)chemical sensing devices

Surface enhanced Raman scattering (SERS) sensors


Yu and White have applied a piezoelectric desktop printer to
deposit aqueous Ag nanoparticle colloids as Raman scattering
enhancers onto paper substrates for a SERS-based sensor
allowing the trace level detection of malathion, heroin, and
cocaine [73, 74]. In contrast to the printing of electrically
conducting traces with Ag nanoparticles, the nanoparticle density required for SERS substrates is significantly lower and,
therefore, allows for the use of diluted water-based printing
inks and standard desktop printers, although requiring multiple
printing cycles (5 or 10 cycles, respectively). During the drying
process, individual nanoparticles aggregate to form silver clusters. In contrast to conventional SERS substrates obtained by
nanofabrication of surfaces with nanoscale metallic structures,
inkjet printing on paper substrates is an extremely low-cost,
simple, and rapid approach. All that is needed for substrate
fabrication is a common desktop printer and an aqueous Agnanoparticle ink, allowing sensor production at the time and
point of use. This not only reduces the costs and time required
to obtain the substrates but also allows overcoming the problem of limited storage stability of Ag-based SERS substrates
due to oxidation. With the simple system developed by the
authors, which includes a lateral flow analyte concentration
step, detection limits as low as 413 pg for malathion, 15 ng for
cocaine, and 9 ng for heroin were achieved.

5799

kept below a certain value. The material inkjet printer used


readily controls both parameters. However, the difficulty to
control random agglomeration of individually deposited droplets into larger pools on the cantilever surface results in nonuniform coverage of the substrate surface after drying. Ness et
al. demonstrated that this disadvantage could be overcome by
applying a droplet swelling method after printing. Glycerol
is widely used as a viscosity modifier in inkjet inks. In addition, its hygroscopic nature makes it simultaneously work as a
humectant, preventing the rapid drying out of ink and the
resulting nozzle clogging. When the cantilevers surfacecoated with a glycerol containing protein ink are kept in a
controlled high-humidity environment for several hours after
printing, inkjet deposited droplets increase in size by uptake of
water, resulting in a uniform fluid coverage of the entire
substrate surface and, after drying, in a more uniform and
dense protein cover layer.
The first example presented in this section is again very
typical for the general tendency in using inkjet printing technology, where low cost, speed and the possibility of mass
production are most important. The second example, however, although probably not very competitive regarding speed of
fabrication, is a good demonstration that inkjet printing is not
only a tool for routine mass fabrication, but continues to be
developed in order to overcome its own limitations.
Microfluidic paper-based analytical devices (PADs)

Sensors with mechanical (mass) transduction


As early as 2004, Bietsch et al. have made use of a piezoelectric inkjet material deposition system for the application of
various active sensing layers (alkanethiol SAM for pH and ion
sensing, single-stranded DNA oligomers for the detection of
complementary DNA, polymer layers for water and ethanol
vapor sensing) on an array of cantilevers [75]. The performance of the SAM coated sensors obtained by inkjet printing
proved to be equivalent to a cantilever array conventionally
modified by insertion into an array of glass capillaries filled
with the coating liquid. The inkjet modified DNA or polymercoated sensors also proved to be fully functional. It was
demonstrated that inkjet printing can replace the capillary
coating technique, resulting in significantly faster sensor fabrication, suitable for automation and large-scale application. In
addition, only inkjet printing allowed for the selective singlesided coating of a cantilever without contaminating the
backside.
In a more recent paper, high spatial control of inkjet printing applied to the coating of cantilevers was nicely demonstrated (Fig. 10) and a method for further enhancement of
deposited layer uniformity has been introduced [76]. In order
to prevent the contamination of the backside by spilling over
of liquid deposited on the topside of a cantilever, droplet size
and the number of droplets deposited per unit area have to be

Considering the historical origin of printing, the combination


of paper with a printing technology such as inkjet seems to be
a very obvious approach. The previous sections have already demonstrated that paper or paper-like materials are
useful and popular substrates for the deposition of active

Fig. 10 Highly precise deposition of biotinylated bovine serum albumin on individual microcantilevers (45 m width, 300 m length) by
piezoelectric inkjet printing. Adapted from Reference [76] Ness SJ,
Kim S, Woolley AT, Nordin GP (2012) Single-sided inkjet
functionalization of silicon photonic microcantilevers. Sens Actuators
B 161:8087, Copyright (2012), with permission from Elsevier

5800

sensing layers by inkjet printing. The microporous structure of


paper allows for the absorption of relatively large amounts of
liquids and for their wicking through the fiber network driven
by capillary forces. These specific characteristics of paper
have given rise to the recent trend to fabricate microfluidic
(bio)analytical devices on paper substrates instead of the more
common glass or polymer platforms, which has been initialized by Whitesides and coworkers in 2007 [77]. As with any
type of microfluidic analytical device, the fabrication of
PADs requires the two principal steps of substrate patterning
and of sensing (bio)chemistry deposition. In the Whitesides
groups original approach, microfluidic patterning was
achieved by photolithography, whereas chemical sensing reagents have been deposited by manual spotting. The field
of PADs and closely related topics has been reviewed
several times [22, 7883]. Here, only a few examples limited
to the application of inkjet printing methods are introduced.
To the best of our knowledge, our research group has been
the first to demonstrate an inkjet printing approach for the
fabrication of PADs. While inkjet technology has been previously used for the deposition of functional materials for
(bio)chemical sensing, we have first demonstrated its applicability to the microfluidic patterning of paper substrates.
Inspired by a report of Schubert and coworkers [14], we have
chosen an approach known as inkjet etching and adapted it
to a filter paper material to create microfluidic channels and
areas for the deposition of (bio)sensing chemistries using a
piezoelectric research-type inkjet printer [15]. In this method,
the printing ink for substrate patterning simply consisted of
toluene, which was inkjet printed onto a filter paper pretreated
by soaking into a solution of hydrophobic poly(styrene)
(Fig. 11a). Toluene ejected from the inkjet nozzle locally
dissolves and displaces the polymer and re-exposes the originally hydrophilic paper, leading to well-defined microfluidic
channels of 420 50 m width. In a second printing step,
aqueous or ethanolic solutions of sensing materials for colorimetric determination of pH (mixture of four chromogenic
pH-indicators), total protein (tetrabromophenol blue) and glucose (glucose oxidase, horseradish peroxidase, o-toluidine,
ascorbic acid) were printed into specific detection areas using
the same inkjet printer, resulting in a PAD as shown in
Fig. 11b. The application of 4.5 L of sample to a central
inlet area was sufficient for the simultaneous semiquantitative
analysis of the three analytes after digital color analysis of
scanned images of the dried PADs.
We have also demonstrated the applicability of the same
inkjet printing approach for the fabrication of a combined
immuno-chemical PAD, where two lateral-flow-type model
immunoassays (mouse IgG, human IgG) and a simple chemical
sensor (pH) were integrated onto a single piece of filter paper
[84]. Inkjet printing was used for the patterning of the
microfluidic structure and the deposition of recognition antibodies required for the immunochromatographic assay. The

N. Komuro et al.

decision to focus on inkjet printing technology in our work is


strongly driven by the possibility to perform both the patterning
of the microfluidic structure and the deposition of sensing
materials by the same method, which results in all inkjet
printed PADs.
Li et al. have first used a standard thermal desktop printer to
fabricate all inkjet printed PADs [85, 86]. Hydrophobic
barriers defining the outline of a microfluidic structure were
directly inkjet printed by using a heptane solution of the paper
sizing agent alkenyl ketene dimer (AKD), which is a common
material for hydrophobization of cellulose applied in paper
making. After printing, patterned substrates were heated at
100 C for 8 min, resulting in the covalent attachment of long
alkyl chains (C16C20) onto cellulose fibers. In the second
inkjet printing process, the reagents for the colorimetric detection of nitrite anions according to the Griess reaction (aqueous solution of citric acid, sulfanilamide and N-(1-naphthyl)ethylenediamine) were deposited on the PAD. The advantages of this method compared with the previously described
approach are that the paper substrate requires no pretreatment
and that heptane is a less aggressive solvent than toluene,
allowing the use of a desktop printer. In addition, the
microfluidic channels and detection areas are not exposed to
any solvents and polymers during the patterning process.
In the following, our research group has reported a different
all inkjet printed approach to PADs using a standard
piezoelectric desktop printer for all fabrication steps [87,
88]. A UV-curable hydrophobic ink composition (octadecyl
acrylate, 1,10-decanediol diacrylate, benzyldimethylketal)
free of volatile organic compounds has been applied for the
direct printing of hydrophobic barriers, and back cover layers
of microfluidic patterns in filter paper. In a double-sided inkjet
printing process, the lines defining the microfluidic structure
are first printed on the topside of the filter paper and cured by
60 s exposure to UV light. Then, a uniform layer of the same
hydrophobic ink is deposited on the backside of the paper and
UV cured for 60 s, resulting in a back cover layer preventing
aqueous samples from penetrating throughout the entire thickness of the paper device (Fig. 12a). Finally, the same inkjet
printer has been used to deposit reagents for chemical sensing
(Fig. 12b), which has been demonstrated on the example of a
simple pH sensor [87] and a hydrogen peroxide sensor [88].
With this patterning method, microfluidic channels as narrow
as 272 19 m have been achieved. This is, to the best of our
knowledge, the narrowest value reported for a standard printing method. For comparison, in their photolithography approach requiring a photomask, the Whitesides group has
fabricated channels of widths down to 186 13 m [83].
All four examples of PADs presented above have been
obtained in an all inkjet printing approach. However, there
are also a number of reports on PADs, where inkjet printing is just used in one part of the entire sensor fabrication
process.

Inkjet printed (bio)chemical sensing devices

5801

Fig. 11 (a) Schematic outline of the fabrication process of an all


inkjet printed PAD featuring microfluidic channels connecting a
central sample inlet area with three different detection areas and a
reference area. Steps 2 (patterning) and 3 (chemical sensing reagent
application) are performed on the same inkjet printing apparatus (the
pen symbol indicates the use of the inkjet printer). (b) Optimized inkjet

printed PAD before sample application. The dotted squares indicating


the sensing areas have been added for illustrative purposes only and
form no part of the sensor pattern. Adapted with permission from
Reference [15] Abe K, Suzuki K, Citterio D (2008) Inkjet printed
microfluidic multianalyte chemical sensing paper. Anal Chem.
80:69286934, Copyright (2008) American Chemical Society

Jayawardane et al. used the microfluidic patterning process developed by Li et al. described above [85] in the
preparation of a low cost PAD for the detection of reactive
phosphate in water [89]. The sensing reagents for the
phosphomolybdenum blue-based phosphate determination
were deposited by manual pipetting.
Delaney et al. relied on the same alkenyl ketene dimer inkjet
printing method for microfluidic patterning to create PADS
for the electrochemiluminescence-based (ECL) sensing of
2-(dibutylamino)-ethanol and nicotinamide adenine dinucleotide (NADH) [90]. The inkjet patterned paper sections were
loaded with a ruthenium(II) complex ECL-transducer by manual pipetting and laminated onto screen printed carbon electrodes to assemble the final sensing devices.
In a report on an all-printed glucose sensor for diagnostics,
Mttnen et al. compared the performance of inkjet printing

and flexographic printing for the patterning of reaction zones


on paper substrates [91]. The material used for the creation of
hydrophobic barriers in paper was a commercially available
vinyl-substituted polydimethylsiloxane. While the ink was directly applicable for flexographic printing, a dilution with xylene was required before inkjet printing with a piezoelectric
research-type printer to reduce the viscosity. Comparable barrier properties were observed for both flexographically and
inkjet printed films. However, in the case of flexographic
printing, six printed layers were necessary to form a hydrophobic barrier throughout the thickness of the paper, while with the
inkjet printing method, a single print cycle was sufficient.
Despite of this difference, the authors concluded that flexographic printing allows for faster patterning with no need to
dilute the commercial ink, whereas inkjet printing results in
better print resolution. For the fabrication of their final all-

Fig. 12 Schematic representation of an inkjet-based PAD fabrication


process: (a) microfluidic patterning of a filter paper by a double-sided
printing process (grey and black colors indicate the printed hydrophobic features before and after UV curing, respectively) using a UV
curable ink consisting of a mixture of hydrophobic monomer and
crosslinker; (b) inkjet printing of sensing ink; all printing steps have

been performed on a standard piezoelectric desktop printer. Reprinted


from Reference [88] RSC Advances, accepted manuscript, Maejima
K.; Tomikawa S.; Suzuki K.; Citterio D.; Inkjet printing: an integrated
and green chemical approach to microfluidic paper-based analytical
devices, Doi: 10.1039/C3RA40828K, Copyright (2013), Reproduced
by permission of The Royal Society of Chemistry

5802

printed colorimetric glucose sensor, the authors have


chosen a combination of flexography for paper patterning (polydimethylsiloxane) and screen printing for sensing chemistry deposition (glucose oxidase, poly(vinyl)
alcohol, potassium iodide, molybdic acid).
Hossain and Brennan demonstrated a PAD for sensitive
detection of various heavy metals based on enzyme inhibition,
where the enzyme, the chromogenic enzyme substrate, and
colorimetric metal ion indicators were inkjet printed (piezoelectric research-type printer) into flow channels and detection
areas of a filter paper, which has been microfluidically patterned using a wax printing device [92].
Fu et al. used a piezoelectric research-type inkjet system in
the patterning of detection antibodies on a two-dimensional
paper-based (lateral-flow through nitrocellulose) immunosensing
device for malaria antigen detection [93]. In this example, the
fluidic pattern has been achieved by cutting out of a larger
substrate sheet (nitrocellulose) by a laser cutter.
Inkjet printing of living cells
Recently, the inkjet printing technology has also attracted
attention as a tool for the fabrication of cellular patterns, with
probably the most fascinating perspective being the creation of
artificial organs [9497]. The use of thermal and piezoelectric
printing technology has been reported, including both conventional desktop printers and specialized 3-D bioprinters. In
terms of analytical applications, inkjet patterning of cells in the
form of microarrays on substrates is a possible approach to
realize biosensors based on living cells, for example for application to drug screening [98]. Inkjet deposited cellular arrays
have also been used for single cell analysis by mass spectrometry [99]. The advantage of inkjet printing lies in the wellcontrolled non-contact deposition of cells to predetermined
locations [100]. A profound discussion of these new developments would go beyond the scope of this work. However, it
should be mentioned that as in the case of biomolecule (e.g.,
enzymes) containing printing inks, the optimization of ink
compositions with the goal to obtain stable printing properties
while retaining the highest degree of cell viability is a key issue
[101, 102]. The influence of jetting parameters (e.g., applied
nozzle voltage) in piezoelectric printing on cell viability has
also been investigated [103]. An additional challenge, although
not directly related to the inkjet printing technology, is the
selection of substrates that prevent deposited cells from drying
and that allow delivery of the nutrients required to maintain
cells alive.

Final remarks
Of course, inkjet printing is not a universally applicable technology. One major limitation comes from the requirement for

N. Komuro et al.

the ink to fit a relatively narrow window of viscosity and


surface tension, both of which are dependent on the printer.
It is not simply enough to prepare the material to be deposited
in a liquid form such as a solution or dispersion, but the
resulting composition must fulfill the viscosity and surface
tension requirements. This automatically results in limits of
the concentration of functional materials in a printing ink,
often leading to the requirement of multiple printing cycles
to achieve the deposition of sufficient amounts of materials.
This requirement may then negatively influence sensing layer
fabrication reproducibility, since small differences in amounts
of repeatedly deposited materials are summing up. In addition,
especially when industrial scale mass production or the sole
use of desktop printers is targeted, the amount of organic
solvents and volatile organic compounds should be reduced.
This creates the challenge of developing inkjet printable liquid
compositions based on aqueous solvent systems alone. In
many cases, it is necessary to use additives to control viscosity
and surface tension, in particular with water as the ink solvent.
Furthermore, inks based on nanoparticles generally require
dispersion stabilizing agents to prevent particle aggregation
and printer nozzle clogging. A simple search in the patent
literature reveals numerous additives that are routinely used in
the formulation of standard color inkjet printing inks.
However, it has to be considered that all nonvolatile additives
(e.g., surfactants, etc.) will remain on the substrate after deposition by inkjet printing. While this does not pose a problem
for standard office printing, for the fabrication of sensing
devices or their components, the presence of materials other
than the active functional sensing compounds is normally not
desired. It is therefore necessary to keep in mind a possible
negative influence of ink additives.
One of the often-mentioned positive characteristics of
inkjet printing is the highly economical use of inks and
the low creation of waste. While this is true in principle, it should be noted that many inkjet systems require
nozzle priming and/or nozzle purging cycles. During these
processes, the amount of wasted ink can be significant.
Furthermore, often a minimal ink amount in the milliliter
order is required because of the dead volume between ink
reservoirs and printer nozzles. These are not serious issues for
large-scale manufacturing, but it becomes relevant during
sensor development and laboratory scale prototyping experiments. One example of overcoming this problem is the replacement of the ink tanks of desktop printers with
micropipette tips to reduce the required minimal volume of
printing ink from the milliliter order down to several tens of
microliters [69].
Despite of remaining challenges in particular related to
printable ink formulation, it has been shown by selected
examples that inkjet printing technology is indeed applicable
for the fabrication of all types of (bio)chemical sensing devices. The motivation for using an inkjet printing system

Inkjet printed (bio)chemical sensing devices

might be different for each specific application. However, the


following list of keywords appears over and over again
throughout the literature and, therefore, best describes the
most valuable characteristics of inkjet printing: simple, low
cost, high speed, precise, reproducible, flexible, non-contact,
low material waste, and applicable to large scale production.
In many cases, inkjet printing has been used to replace an
alternative material deposition method, and often improvements in fabrication reproducibility or sensor performance
have been reported. In other situations, no such improvements
were achieved, but the advantages gained in production speed,
simplicity, and cost still favor the inkjet-based approach.
There are also some characteristics of inkjet technology that
go beyond that list of keywords. For example, the nature of
inkjet printing, where materials are deposited in the form of
single small-sized droplets in contrast to a continuous layer,
allows not only to control the thickness of an applied layer
(e.g., multiple print cycles), but also the density of the deposited structure (e.g., variations in drop spacing). The variation
of these two parameters can, in some cases, result in sensing
layers with different physical and chemical properties, although an identical liquid has been deposited.
The number of (bio)chemical sensors fabricated in research
laboratories with no other equipment but an off-the-shelf
desktop printer will certainly continue to increase. At the same
time, (bio)chemical sensing devices can be expected to become more widespread and economical, once their inkjetbased manufacturing approach is hopefully transferred to
industrial scale mass production.

References
1. Kimura J, Kawana Y, Kuriyama T (1988) An immobilized enzyme membrane fabrication method using an ink jet nozzle.
Biosensors 4:4152
2. Sirringhaus H, Kawase T, Friend RH, Shimoda T, Inbasekaran M,
Wu W, Woo EP (2000) High-resolution inkjet printing of allpolymer transistor circuits. Science 290:2122126
3. Calvert P (2001) Inkjet printing for materials and devices. Chem
Mater 13:32993305
4. De Gans BJ, Duineveld PC, Schubert US (2004) Inkjet printing
of polymers: State of the art and future developments. Adv Mater
16:203213
5. Tekin E, Smith PJ, Schubert US (2008) Inkjet printing as a
deposition and patterning tool for polymers and inorganic particles. Soft Matter 4:703713
6. Singh M, Haverinen HM, Dhagat P, Jabbour GE (2010) Inkjet
printingprocess and its applications. Adv Mater 22:673685
7. Hwang SY, Lim G (2000) DNA chip technologies. Biotechnol
Bioprocess Eng 5:159163
8. Gonzalez-Macia L, Morrin A, Smyth MR, Killard AJ (2010)
Advanced printing and deposition methodologies for the fabrication of biosensors and biodevices. Analyst 35:845867
9. Delaney JT, Smith PJ, Schubert US (2009) Inkjet printing of
proteins. Soft Matter 5:48664877

5803
10. Kukkola J, Mohl M, Leino AR, Toth G, Wu MC, Shchukarev A,
Popov A, Mikkola JP, Lauri J, Riihimaki M, Lappalainen J,
Jantunen H, Kordas K (2012) Inkjet printed gas sensors: metal
decorated WO3 nanoparticles and their gas sensing properties. J
Mater Chem 22:1787817886
11. Teichler A, Perelaer J, Schubert US (2013) Inkjet printing of
organic electronics - comparison of deposition techniques and
state-of-the-art developments. J Mater Chem C 1:19101925
12. Barbulovic-Nad I, Lucente M, Yu S, Mingjun Z, Wheeler AR,
Bussmann M (2006) Bio-microarray fabrication techniques - A
review. Crit Rev Biotechnol 26:237259
13. Metters JP, Kadara RO, Banks CE (2011) New directions in
screen printed electroanalytical sensors: an overview of recent
developments. Analyst 136:10671076
14. De Gans BJ, Hoeppener S, Schubert US (2006) Polymer-relief
microstructures by inkjet etching. Adv Mater 18:910914
15. Abe K, Suzuki K, Citterio D (2008) Inkjet printed microfluidic
multianalyte chemical sensing paper. Anal Chem 80:69286934
16. Nie Z, Kumacheva E (2008) Patterning surfaces with functional
polymers. Nat Mater 7:277290
17. Xia Y, Whitesides GM (1998) Soft lithography. Angew Chem, Int
Ed 37:550575
18. Piner RD, Zhu J, Xu F, Hong S, Mirkin CA (1999) "Dip-pen"
nanolithography. Science 283:661663
19. Derby B (2008) Bioprinting: inkjet printing proteins and hybrid
cell-containing materials and structures. J Mater Chem 18:5717
5721
20. Setti L, Fraleoni-Morgera A, Ballarin B, Filippini A, Frascaro D,
Piana C (2005) An amperometric glucose biosensor prototype
fabricated by thermal inkjet printing. Biosens Bioelectron
20:20192026
21. Setti L, Fraleoni-Morgera A, Mencarelli I, Filippini A, Ballarin B,
Di Biase M (2007) An HRP-based amperometric biosensor fabricated by thermal inkjet printing. Sens Actuators, B 126:252
257
22. Di Risio S, Yan N (2010) Bioactive paper through inkjet printing.
J Adhes Sci Technol 24:661684
23. Khan MS, Fon D, Li X, Tian J, Forsythe J, Garnier G, Shen W
(2010) Biosurface engineering through ink jet printing. Colloids
Surf B 75:441447
24. Nishioka GM, Markey AA, Holloway CK (2004) Protein damage
in drop-on-demand printers. J Am Chem Soc 126:1632016321
25. Di Risio SD, Yan N (2007) Piezoelectric ink-jet printing of
horseradish peroxidase: Effect of Ink viscosity modifiers on
activity. Macromol Rapid Commun 28:19341940
26. Loffredo F, Mauro ADGD, Burrasca G, La Ferrara V, Quercia L,
Massera E, Di Francia G, Sala DD (2009) Ink-jet printing technique in polymer/carbon black sensing device fabrication. Sens
Actuators, B 143:421429
27. Ihalainen P, Majumdar H, Mttnen A, Wang S, sterbacka R,
Peltonen J (2012) Versatile characterization of thiol-functionalized
printed metal electrodes on flexible substrates for cheap diagnostic
applications. Biochim Biophys Acta - Gen Subjects. doi:10.1016/
j.bbagen.2012.09.007
28. Jalkanen T, Mkil E, Mttnen A, Tuura J, Kaasalainen M,
Lehto V-P, Ihalainen P, Peltonen J, Salonen J (2012) Porous
silicon micro- and nanoparticles for printed humidity sensors.
Appl Phys Lett 101:263110263114
29. Andersson H, Manuilskiy A, Unander T, Lidenmark C, Forsberg
S, Nilsson HE (2012) Inkjet printed silver nanoparticle humidity
sensor with memory effect on paper. IEEE Sens J 12:19011905
30. Sarfraz J, Tobjrk D, sterbacka R, Lindn M (2012) Low-cost
hydrogen sulfide gas sensor on paper substrates: Fabrication and
demonstration. IEEE Sens J 12:19731978
31. Mttnen A, Vanamo U, Ihalainen P, Pulkkinen P, Tenhu H,
Bobacka J, Peltonen J (2013) A low-cost paper-based inkjet-

5804

32.

33.

34.

35.

36.

37.

38.

39.
40.

41.

42.

43.

44.

45.

46.

47.
48.

49.

N. Komuro et al.
printed platform for electrochemical analyses. Sens Actuators, B
177:153162
Altenberend U, Molina-Lopez F, Oprea A, Briand D, Brsan N, De
Rooij NF, Weimar U (2012) Towards fully printed capacitive gas
sensors on flexible PET substrates based on Ag interdigitated transducers with increased stability. Sens Actuators, B. doi:10.1016/
j.snb.2012.11.025
Claramunt S, Monereo O, Boix M, Leghrib R, Prades JD, Cornet
A, Merino P, Merino C, Cirera A (2013) Flexible gas sensor array
with an embedded heater based on metal decorated carbon
nanofibres. Sens Actuators, B. doi:10.1016/j.snb.2012.12.093
Bidoki SM, Lewis DM, Clark M, Vakorov A, Millner PA,
McGorman D (2007) Ink-jet fabrication of electronic components. J Micromech Microeng 17:967
Magdassi S, Grouchko M, Berezin O, Kamyshny A (2010)
Triggering the sintering of silver nanoparticles at room temperature. ACS Nano 4:19431948
Perelaer J, Abbel R, Wnscher S, Jani R, van Lammeren T,
Schubert US (2012) Roll-to-roll compatible sintering of inkjet
printed features by photonic and microwave exposure: From nonconductive ink to 40% bulk silver conductivity in less than 15
seconds. Adv Mater 24:26202625
Tobjrk D, Aarnio H, Pulkkinen P, Bollstrm R, Mttnen A,
Ihalainen P, Mkel T, Peltonen J, Toivakka M, Tenhu H,
sterbacka R (2012) IR-sintering of ink-jet printed metalnanoparticles on paper. Thin Solid Films 520:29492955
Mttnen A, Ihalainen P, Pulkkinen P, Wang S, Tenhu HJ,
Peltonen J (2012) Inkjet-printed gold electrodes on paper - characterisation and functionalisation. ACS Appl Mater Interfaces
4:955964
Nie X, Wang H, Zou J (2012) Inkjet printing of silver citrate
conductive ink on PET substrate. Appl Surf Sci 261:554560
Ihalainen P, Majumdar H, Viitala T, Trngren B, Nrjeoja T,
Mttnen A, Sarfraz J, Hrm H, Yliperttula M, sterbacka R,
Peltonen J (2012) Application of paper-supported printed gold
electrodes for impedimetric immunosensor development.
Biosensors 3:117
Jensen GC, Krause CE, Sotzing GA, Rusling JF (2011) Inkjetprinted gold nanoparticle electrochemical arrays on plastic.
Application to immunodetection of a cancer biomarker protein.
Phys Chem Chem Phys 13:48884894
Hu C, Bai X, Wang Y, Jin W, Zhang X, Hu S (2012) Inkjet
printing of nanoporous gold electrode arrays on cellulose membranes for high-sensitive paper-like electrochemical oxygen sensors using ionic liquid electrolytes. Anal Chem 84:37453750
Heller A, Feldman B (2008) Electrochemical glucose sensors and
their applications in diabetes management. Chem Rev 108:2482
2505
Kords K, Mustonen T, Tth G, Jantunen H, Lajunen M, Soldano
C, Talapatra S, Kar S, Vajtai R, Ajayan PM (2006) Inkjet printing
of electrically conductive patterns of carbon nanotubes. Small
2:10211025
Small WR, in het Panhuis M (2007) Inkjet printing of transparent,
electrically conducting single-walled carbon-nanotube composites. Small 3:15001503
Dua V, Surwade SP, Ammu S, Agnihotra SR, Jain S, Roberts KE,
Park S, Ruoff RS, Manohar SK (2010) All-organic vapor sensor
using inkjet-printed reduced graphene oxide. Angew Chem, Int
Ed 49:21542157
Weng B, Shepherd RL, Crowley K, Killard AJ, Wallace GG
(2010) Printing conducting polymers. Analyst 135:27792789
Su S-J, Kuramoto N (2000) Synthesis of processable polyaniline
complexed with anionic surfactant and its conducting blends in
aqueous and organic system. Synth Met 108:121126
Crowley K, Morrin A, Hernandez A, O'Malley E, Whitten PG,
Wallace GG, Smyth MR, Killard AJ (2008) Fabrication of an

ammonia gas sensor using inkjet-printed polyaniline nanoparticles.


Talanta 77:710717
50. Crowley K, O'Malley E, Morrin A, Smyth MR, Killard AJ (2008)
An aqueous ammonia sensor based on an inkjet-printed polyaniline
nanoparticle-modified electrode. Analyst 133:391399
51. Morrin A, Ngamna O, O'Malley E, Kent N, Moulton SE, Wallace
GG, Smyth MR, Killard AJ (2008) The fabrication and characterization of inkjet-printed polyaniline nanoparticle films.
Electrochim Acta 53:50925099
52. Oh W-K, Kim S, Shin K-H, Jang Y, Choi M, Jang J (2013) Inkjetprinted polyaniline patterns for exocytosed molecule detection
from live cells. Talanta 105:333339
53. Phongphut A, Sriprachuabwong C, Wisitsoraat A,
Tuantranont A, Prichanont S, Sritongkham P (2013) A disposable amperometric biosensor based on inkjet-printed Au/
PEDOT-PSS nanocomposite for triglyceride determination.
Sens Actuators, B 178:501507
54. Sriprachuabwong C, Karuwan C, Wisitsorrat A, Phokharatkul D,
Lomas T, Sritongkham P, Tuantranont A (2012) Inkjet-printed
graphene-PEDOT:PSS modified screen printed carbon electrode
for biochemical sensing. J Mater Chem 22:54785485
55. Karuwan C, Sriprachuabwong C, Wisitsoraat A, Phokharatkul D,
Sritongkham P, Tuantranont A (2012) Inkjet-printed graphenepoly(3,4-ethylenedioxythiophene):poly(styrene-sulfonate) modified on screen printed carbon electrode for electrochemical sensing of salbutamol. Sens Actuators, B 161:549555
56. Tseng CC, Chou YH, Hsieh TW, Wang MW, Shu YY, Ger MD
(2012) Interdigitated electrode fabricated by integration of ink-jet
printing with electroless plating and its application in gas sensor.
Colloid Surface A 402:4552
57. Mabrook MF, Pearson C, Petty MC (2006) Inkjet-printed polymer films for the detection of organic vapors. IEEE Sens J
6:14351444
58. Yang M, Li LH, Zhang SQ, Li GY, Zhao HJ (2010) Preparation,
characterisation and sensing application of inkjet-printed nanostructured TiO2 photoanode. Sens Actuators, B 147:622628
59. O'Toole M, Shepherd R, Wallace GG, Diamond D (2009) Inkjet
printed LED based pH chemical sensor for gas sensing. Anal
Chim Acta 652:308314
60. Courbat J, Briand D, Wollenstein J, de Rooij NF (2011) Polymeric
foil optical waveguide with inkjet printed gas sensitive film for
colorimetric sensing. Sens Actuators, B 160:910915
61. Courbat J, Briand D, Damon-Lacoste J, Wllenstein J, de Rooij
NF (2009) Evaluation of pH indicator-based colorimetric films
for ammonia detection using optical waveguides. Sens Actuators,
B 143:6270
62. Yoon B, Ham DY, Yarimaga O, An H, Lee CW, Kim JM (2011)
Inkjet printing of conjugated polymer precursors on paper substrates for colorimetric sensing and flexible electrothermochromic
display. Adv Mater 23:54925497
63. Yoon B, Shin H, Yarimaga O, Ham DY, Kim J, Park IS, Kim JM
(2012) An inkjet-printable microemulsion system for colorimetric
polydiacetylene supramolecules on paper substrates. J Mater
Chem 22:86808686
64. Alkasir RSJ, Ornatska M, Andreescu S (2012) Colorimetric paper
bioassay for the detection of phenolic compounds. Anal Chem
84:97299737
65. Hossain SMZ, Luckham RE, Smith AM, Lebert JM, Davies LM,
Pelton RH, Filipe CDM, Brennan JD (2009) Development of a
bioactive paper sensor for detection of neurotoxins using piezoelectric inkjet printing of sol-gel-derived bioinks. Anal Chem
81:54745483
66. Hossain SMZ, Luckham RE, McFadden MJ, Brennan JD (2009)
Reagentless bidirectional lateral flow bioactive paper sensors for
detection of pesticides in beverage and food samples. Anal Chem
81:90559064

Inkjet printed (bio)chemical sensing devices


67. Hossain S, Ozimok C, Sicard C, Aguirre S, Ali M, Li Y, Brennan
J (2012) Multiplexed paper test strip for quantitative bacterial
detection. Anal Bioanal Chem 403:15671576
68. Kwon H, Samain F, Kool ET (2012) Fluorescent DNAs printed
on paper: Sensing food spoilage and ripening in the vapor phase.
Chem Sci 3:25422549
69. Yu A, Shang J, Cheng F, Paik BA, Kaplan JM, Andrade RB,
Ratner DM (2012) Biofunctional paper via the covalent modification of cellulose. Langmuir 28:1126511273
70. Carter JC, Alvis RM, Brown SB, Langry KC, Wilson TS,
McBride MT, Myrick ML, Cox WR, Grove ME, Colston BW
(2006) Fabricating optical fiber imaging sensors using inkjet
printing technology: A pH sensor proof-of-concept. Biosens
Bioelectron 21:13591364
71. Gervais L, Delamarche E (2009) Toward one-step point-of-care
immunodiagnostics using capillary-driven microfluidics and
PDMS substrates. Lab Chip 9:33303337
72. Kirk JT, Fridley GE, Chamberlain JW, Christensen ED, Hochberg
M, Ratner DM (2011) Multiplexed inkjet functionalization of
silicon photonic biosensors. Lab Chip 11:13721377
73. Yu WW, White IM (2010) Inkjet printed surface enhanced raman
spectroscopy array on cellulose paper. Anal Chem 82:96269630
74. Yu WW, White IM (2013) Inkjet-printed paper-based SERS dipsticks
and swabs for trace chemical detection. Analyst 138:10201025
75. Bietsch A, Zhang JY, Hegner M, Lang HP, Gerber C (2004)
Rapid functionalization of cantilever array sensors by inkjet
printing. Nanotechnology 15:873880
76. Ness SJ, Kim S, Woolley AT, Nordin GP (2012) Single-sided
inkjet functionalization of silicon photonic microcantilevers. Sens
Actuators, B 161:8087
77. Martinez AW, Phillips ST, Butte MJ, Whitesides GM (2007)
Patterned paper as a platform for inexpensive, low-volume, portable bioassays. Angew Chem, Int Ed 46:13181320
78. Ballerini D, Li X, Shen W (2012) Patterned paper and alternative
materials as substrates for low-cost microfluidic diagnostics.
Microfluid Nanofluidics 13:769787
79. Parolo C, Merkoci A (2013) Paper-based nanobiosensors for
diagnostics. Chem Soc Rev 42:450457
80. Li X, Ballerini DR, Shen W (2012) A perspective on paper-based
microfluidics: Current status and future trends. Biomicrofluidics
6:011301011313
81. Martinez AW (2011) Microfluidic paper-based analytical devices:
from POCKET to paper-based ELISA. Bioanalysis 3:25892592
82. Pelton R (2009) Bioactive paper provides a low-cost platform for
diagnostics. TrAC, Trends Anal Chem 28:925942
83. Martinez AW, Phillips ST, Whitesides GM, Carrilho E (2010)
Diagnostics for the developing world: Microfluidic paper-based
analytical devices. Anal Chem 82:310
84. Abe K, Kotera K, Suzuki K, Citterio D (2010) Inkjet-printed
paperfluidic immuno-chemical sensing device. Anal Bioanal
Chem 398:885893
85. Li X, Tian J, Garnier G, Shen W (2010) Fabrication of paper-based
microfluidic sensors by printing. Colloids Surf B 76:564570
86. Li X, Tian J, Shen W (2010) Progress in patterned paper sizing
for fabrication of paper-based microfluidic sensors. Cellulose
17:649659
87. Citterio D, Maejima K, Suzuki K (2011) VOC-free inkjet patterning method for the fabrication of "paperfluidic" sensing devices. Paper presented at the 15th International Conference on
Miniaturized Systems for Chemistry and Life Sciences, Seattle,
Washington, USA
88. Maejima K, Tomikawa S, Suzuki K, Citterio D (2013) Inkjet
printing: An integrated and green chemical approach to

5805

89.

90.

91.

92.

93.

94.
95.
96.

97.
98.

99.

100.

101.

102.

103.

104.

105.

106.

107.

108.

microfluidic paper-based analytical devices. RSC Adv.


doi:10.1039/C3RA40828K
Jayawardane BM, McKelvie ID, Kolev SD (2012) A paper-based
device for measurement of reactive phosphate in water. Talanta
100:454460
Delaney JL, Hogan CF, Tian JF, Shen W (2011) Electrogenerated
chemiluminescence detection in paper-based microfluidic sensors. Anal Chem 83:13001306
Mttnen A, Fors D, Wang S, Valtakari D, Ihalainen P, Peltonen
J (2011) Paper-based planar reaction arrays for printed diagnostics. Sens Actuators, B 160:14041412
Hossain SMZ, Brennan JD (2011) -Galactosidase-based colorimetric paper sensor for determination of heavy metals. Anal
Chem 83:87728778
Fu E, Liang T, Spicar-Mihalic P, Houghtaling J, Ramachandran S,
Yager P (2012) Two-dimensional paper network format that enables
simple multistep assays for use in low-resource settings in the
context of malaria antigen detection. Anal Chem 84:45744579
Calvert P (2007) Printing cells. Science 318:208209
Xu T, Jin J, Gregory C, Hickman JJ, Boland T (2005) Inkjet
printing of viable mammalian cells. Biomaterials 26:9399
Nakamura M, Kobayashi A, Takagi F, Watanabe A, Hiruma Y,
Ohuchi K, Iwasaki Y, Horie M, Morita I, Takatani S (2005)
Biocompatible inkjet printing technique for designed seeding of
individual living cells. Tissue Eng 11:16581666
Derby B (2012) Printing and prototyping of tissues and scaffolds.
Science 338:921926
Feng X, JinHui W, ShuQi W, Naside Gozde D, Umut Atakan G,
Utkan D (2011) Microengineering methods for cell-based
microarrays and high-throughput drug-screening applications.
Biofabrication 3:034101
Ellis SR, Ferris CJ, Gilmore KJ, Mitchell TW, Blanksby SJ, in het
Panhuis M (2012) Direct lipid profiling of single cells from inkjet
printed microarrays. Anal Chem 84:96799683
Liberski AR, Delaney JT, Schubert US (2010) One cellone
well: a new approach to inkjet printing single cell microarrays.
ACS Comb Sci 13:190195
Ferris CJ, Gilmore KJ, Beirne S, McCallum D, Wallace GG, in
het Panhuis M (2013) Bio-ink for on-demand printing of living
cells. Biomater Sci 1:224230
Shabnam P, Madhuja G, Frdric L, Karen CC (2010) Effects of
surfactant and gentle agitation on inkjet dispensing of living cells.
Biofabrication 2:025003
Saunders RE, Gough JE, Derby B (2008) Delivery of human
fibroblast cells by piezoelectric drop-on-demand inkjet printing.
Biomaterials 29:193203
Kit-Anan W, Olarnwanich A, Sriprachuabwong C, Kuruwan C,
Tuantranont A, Wisitsoraat A, Srituravanich W, Pimpin A (2012)
Disposable paper-based electrochemical sensor utilizing inkjetprinted polyaniline modified screen-printed carbon electrode for
ascorbic acid detection. J Electroanal Chem 685:7278
Su SX, Ali M, Filipe CDM, Li YF, Pelton R (2008) Microgel-based
inks for paper-supported biosensing applications. Biomacromolecules
9:935941
Lu Y, Shi W, Jiang L, Qin J, Lin B (2009) Rapid prototyping of
paper-based microfluidics with wax for low-cost, portable bioassay. Electrophoresis 30:14971500
Jung W, Sahner K, Leung A, Tuller HL (2009) Acoustic wavebased NO2 sensor: Ink-jet printed active layer. Sens Actuators, B
141:485490
Shen W, Li M, Ye C, Jiang L, Song Y (2012) Direct-writing
colloidal photonic crystal microfluidic chips by inkjet printing for
label-free protein detection. Lab Chip 12:30893095

Вам также может понравиться