Microarray Bioprinting Technology
Microarray Bioprinting Technology
Microarray
Bioprinting
Technology
Fundamentals and Practices
Microarray Bioprinting Technology
Moo-Yeal Lee
Microarray Bioprinting
Technology
Fundamentals and Practices
Moo-Yeal Lee
Department of Chemical and Biomedical Engineering
Cleveland State University
Washkewicz College of Engineering
Cleveland, OH, USA
v
vi Preface
impossible for me to complete this book. I hope that this book will serve as a valu-
able reference manual for graduate students, postdoctoral fellows, scientists, and
researchers working in this area of research. We envision that microarray bioprint-
ing technology offers new opportunities for creating highly organized multicellular
tissue constructs by precisely dispensing multiple human cell types in hydrogel lay-
ers on a chip with printing robots and mimicking the microenvironments of tissues
in vivo, thereby potentially revolutionizing regenerative medicine, oncology, and
drug discovery.
vii
viii Contents
xiii
Chapter 1
Overview of Microarray Bioprinting
Technology
Moo-Yeal Lee
Contents
1.1 Overview............................................................................................................................ 1
1.2 Microarray Chip Platforms................................................................................................ 3
1.2.1 Chips and Instrument............................................................................................. 3
1.2.2 Cell Microarrays.................................................................................................... 6
1.2.3 Enzyme Microarrays.............................................................................................. 7
1.2.4 Virus Microarrays.................................................................................................. 11
1.3 Experimental Procedures................................................................................................... 12
1.3.1 Cell Spotting.......................................................................................................... 13
1.3.2 Enzyme and Compound Spotting.......................................................................... 13
1.3.3 Chip Incubation, Staining, Scanning, and Data Analysis...................................... 13
1.4 Advantages and Applications............................................................................................ 14
References................................................................................................................................... 16
1.1 Overview
Fig. 1.1 Schematics of (A) a micropillar chip and (B) a microwell chip. (C) Image of micropillar
and microwell chips to scale against a microscopic glass slide.
4 1 Overview of Microarray Bioprinting Technology
Syringe
pumps
Chilling
chip deck
-
Chip-loading deck
Mercury lamp
Objective lens
Color CCD
Solenoid camera
valves Ceramic tips
96-Well plate
4 Channel filter wheel
(A) Water bath Sonicator (B)
Fig. 1.2 Photographs of (A) a microarray spotter (S+ MicroArrayer) with six solenoid valves and
ceramic tips for printing biological samples and (B) a chip-scanning system (S+ Scanner) for cell
image acquisition. The microarray spotter has a printing head with six solenoid valves and a chill-
ing capability that is designed for printing temperature-sensitive hydrogels such as Matrigel.
Several human cell types can be spotted in hydrogels such as alginate, fibrinogen, PuraMatrix, and
Matrigel onto micropillars (typically 60 nL), and various combinations of biological samples as
well as test compounds (typically 950 nL) can be dispensed into microwells. Two chips can be
sandwiched together or separated each other consistently and accurately. The sandwiched chips
can be incubated in a gas-permeable chamber for cell growth.
DataChips
Staining dye
(A)
Sandwiched chips (B)
Fig. 1.3 (A) Gas-permeable incubation chamber and (B) Staining plate. The sandwiched chips are
incubated in a chamber that allows good gas exchange. Cell growth occurs during incubation due
to uniform gas exchange over the entire chip. The micropillar chips with cells (DataChip) can be
stained uniformly on a staining plate for various cell-based assays.
assays can be performed on the chip platform (Fig. 1.3). For complete automa-
tion, a microarray spotter has been developed to enable consistent manipulation
of the micropillar/microwell chip, and a chip-scanning system and image analy-
sis software have been developed to analyze data from the scanned images in a
high-throughput fashion (Figs. 1.2 and 1.4).
1.2 Microarray Chip Platforms 5
Fig. 1.4 S+ Chip Analysis from Samsung for cell image analysis. S+ Chip Analysis extracts
green and red fluorescent intensities from living and dead cells in each cell spot on the scanned
chip, plots sigmoidal dose-response curves with the percentage of live cells against the concentra-
tion of the test compound, and then calculates IC50 values for each test condition. It takes about
10 min to process three DataChips, which is equivalent to seventy-two individual dose response
curves and IC50 values.
Hydrogel
bottom
Fig. 1.6 The DataChip—A human 3D cell culture microarray platform: (A) Microscopic pictures
of the sandwiched chips (left) and live Hep3B human hepatoma cell line encapsulated on the
micropillar chip (right), (B) Scanned image of the entire micropillar chip with live Hep3B cells
stained with a green fluorescent dye, (C) Culture of Hep3B cells on the microwell chip over time,
forming 3D spheroids.
The MetaChip with active human (or rat) metabolizing enzymes (mixtures or human
liver microsomes) encapsulated in a hydrogel matrix has been developed to investi-
gate drug metabolism and enzyme inhibition by compounds (Fig. 1.7). Typical
DMEs printed in the MetaChip include cytochrome P450 (CYP450), UDP-
glycosyltransferase (UGT), sulfotransferase (SULT), and glutathione S-transferase
(GST), among others (Table 1.3). The MetaChip allows the production and screen-
ing of human (or animal) metabolites on the chip platform because enzymatic
kinetic parameters are comparable with conventional approaches. Since it relies on
nanoliter-scale enzyme printing, the approach can save the expense of reagents and
test compounds significantly (ca. 100-fold reduction in assay volume compared to
the 96-well plate counterpart).
8 1 Overview of Microarray Bioprinting Technology
Compounds
(720 nL)
Metabolizing enzymes
(120 nL)
Fig. 1.7 Schematics of the MetaChip for drug metabolism and inhibition.
Table 1.3 Breadth of drug metabolizing enzymes (DMEs) used on the MetaChip
Drug metabolizing enzymes Name
Human Phase I enzymes Cytochrome P450 (CYP1A2, 2C9, 2C19, 2D6, 2E1, 3A4, 3A5)
Flavin monooxygenase (FMO1, FMO3, FMO5)
Monoamine oxidase (MAO-A, MAO-B)
Others (Myeloperoxidase, Esterase, Epoxide hydrolase, DT
diaphorase, NQQ1)
Human Phase II enzymes UDP-glycosyltransferase (UGT1A1, 1A4, 2B4, 2B7)
Sulfotransferase (SULT1A1, 1A3, 1B1)
Glutathione S-transferase (GST)
N-Acetyltransferase (NAT1 and NAT2)
Pooled human liver enzymes Human liver microsomes (HLM)
Human s9 fractions (Hs9)
Rat Phase I enzymes Suite of rat CYP450s
Pooled rat liver enzymes Rat liver microsomes (RLM)
Rat s9 fractions (Rs9)
Fig. 1.8 Scanned image of the microwell chip with varying concentrations of fluorescent dyes
obtained by Cellomics ToxInsight In Vitro Toxicity (IVT) platform.
with the DataChip, the MetaChip has been used for analyzing metabolism-induced
toxicity. For example, the DataChip with Hep3B cells was stamped onto the MetaChip
containing metabolic enzymes (no enzyme, P450 Mix, All Mix, and HLMs) and com-
pounds (six compounds at six different dosages per compound) [3, 7, 9]. No enzyme
contains baculosome® negative control, P450 Mix is a mixture of major P450 iso-
forms, All Mix is a mixture of P450 Mix and Phase II enzymes, and HLMs are pooled
human liver microsomes. This approach allows one to measure not only the toxicity
of parent compounds but also the augmented toxicity by drug metabolites, which are
critically important to predict human drug metabolism in the livers after drug admin-
istration. Thus, it enables early detection of hidden toxicity of metabolites during lead
generation, lead optimization, and preclinical phases of drug development.
The Multizyme Chip, the variant of the MetaChip, encompasses microarrays
of biosynthetic enzymes on a chip to construct artificial synthetic pathways
(Fig. 1.9). Biosynthetically important enzymes are encapsulated into high-density
microarray spots that have the capacity to perform large numbers of enzymatic
reactions on a single chip, ultimately generating a large number of derivatives
starting from a lead compound. Thus, the Multizyme Chip is specifically designed
10 1 Overview of Microarray Bioprinting Technology
Lead compound
(100 nL)
Synthetic enzymes
(150 nL)
Recombinant
adenoviruses
(800 nL)
Fig. 1.11 The scanned image of the DataChip containing THLE-2 cells after sandwiching with
the TeamChip containing recombinant adenoviruses carrying GFP/RFP genes (Ad-GFP/RFP).
Human cells (e.g., THLE-2, Hep3B, HepG2, and Beas-2b) in Matrigel spots were 100 % infected
at 60 multiplicity of infection (MOI).
12 1 Overview of Microarray Bioprinting Technology
To attach cell spots on the micropillar chip, a mixture of poly-L-lysine (PLL) and
barium chloride (BaCl2) is prepared by mixing a 1:2 volume ratio of 0.01 % (wt/vol)
PLL and 24 mM BaCl2. The DataChip is prepared by spotting 60 nL of the PLL/
BaCl2 mixture onto 532 micropillars followed by printing 60 nL of hepatic cell
suspension in alginate on top of the dried PLL/BaCl2 spots. While printing cells, the
micropillar chip is placed on a chilling chip deck at 4 °C to retard evaporation of
water in the spots. A suspension of cells in low-viscosity alginate is prepared by
mixing cell suspension in FBS-supplemented growth medium with alginate solution
in distilled water so that the final concentration of cells and alginate are 2–6 × 106
cells per milliliter and 0.75–1 % (wt/vol), respectively. After nearly instantaneous
gelation, cell spots on the 532 micropillars are immersed in growth medium in the
complementary 532 microwells by stamping. The sandwiched DataChip is incu-
bated in a CO2 incubator at 37 °C for 18 h prior to toxicity tests.
The MetaChip is transversely divided into typically four regions (A–D) by printing
120 nL of metabolic enzyme/Matrigel mixtures in the microwell chip laid on a chilling
slide deck at 4 °C. Specifically, regions A to D contain no enzyme as a test compound
only control, a mixture of human cytochrome P450 isoforms (P450 Mix), a mixture of
P450 Mix and human Phase II metabolizing enzymes (All Mix), and human liver
microsomes (HLM). Immediately after spotting, the MetaChip is stored in a −80 °C
freezer until use. For toxicity tests, six different compounds are printed in regions 1–6
of the MetaChip, creating twenty-four distinct regions on the chip, each region contain-
ing a 3 × 6 mini-array. Within each mini-array, six different doses of a compound are
assayed for toxicity. Thus, a single Data chip combined with a single MetaChip has the
capability to generate twenty-four dose response curves for six compounds and their
metabolites generated from three different metabolic conditions on the chip.
After stamping the DataChip onto the MetaChip and incubating the combined chips
for 24 h in the presence of compounds followed by 48 h post-stamping incubation
in fresh growth medium, the DataChip containing cells is stained with a live/dead
viability/cytotoxicity kit from Life Technologies. Staining dyes such as calcein AM
and ethidium homodimer-1 are used to produce a green fluorescent response from
live cells and a red fluorescent signal from dead cells. The dried DataChip is scanned
with S+ Scanner to obtain fluorescent images of cell spots. Alternatively, a GenePix
Professional 4200 A scanner and Cellomics ToxInsight In Vitro Toxicity (IVT)
14 1 Overview of Microarray Bioprinting Technology
Fig. 1.13 (A) GenePix 4200 A microarray scanner from Molecular Devices and (B) Cellomics
ToxInsight In Vitro Toxicity (IVT) platform from Thermo Fisher are required to obtain scanned
images from the DataChip/MetaChip.
Table 1.5 Comparison of the DataChip/MetaChip platform with the 96-well plate platform
Compound Compound
solution solution/cell Incubation
Platform Spot volume Morphology volume ratio (nL/cell) time (h)
DataChip 60 nL (200 cells/ 3D cells in – – 24–72
(532-micropillar spot) alginate/
chip) Matrigel
MetaChip 120 nL (54 nM Metabolizing 720 nL 3.6 (720 nL/200 24
(532-microwell enzyme) enzymes cells)
chip) in alginate/
Matrigel
Well plate (96-well 100 μL (10,000 2D cell 100 μL 10 (100,000 24–72
plate) cells/well) monolayers nL/10,000 cells)
platform can be used (Fig. 1.13). The dose response curves and IC50 values are
c alculated from scanned images using ImageJ and S+ Chip Analysis.
The fundamentally distinctive advantages of the DataChip/MetaChip platform
are the result of product and process characteristics designed and engineered to
replicate the human metabolism process. Consequently, the platform features a 3D
cellular environment that mimics that of human tissues in vivo. It has the ability to
accommodate a diverse array of cells, metabolizing enzymes and enzyme mixtures,
and a broad array of drug candidates. Combined with platform aspects that optimize
cell growth, these system characteristics transform to robust competitive strengths
of high throughput, high predictive reliability, low cost, and consistent reproduc-
ibility of toxicity profiles (Table 1.5).
The microarray chip technology offers several clear advantages over more conven-
tional in vitro toxicology screening tools. Specifically, it requires extremely small
amounts of enzymes, viruses, cells, compounds, and reagents for analysis. It is well
1.4 Advantages and Applications 15
suited for early stage HTS of compound libraries, which include drug candidates,
chemicals, cosmetic ingredients, and environmental toxicants. The cell encapsula-
tion technology developed is flexible and allows one to culture multiple cell types
from different organs in hydrogel droplets on the chip, thus providing more predic-
tive insight into potential organ-specific toxicity of compounds. 3D-cell cultures on
the chip may provide an environment that simulates the in vivo ECM conditions,
and therefore help to maintain the specific biochemical functions and morphological
features of human cells similar to those found in human organ tissues. Through the
use of recombinant adenoviruses, the chip enables the controlled expression of each
human metabolizing enzyme as well as various combinations of multiple enzymes
in human cells. The gene transduction technology on the chip can be extended to
other cell-based in vitro assays, including gain- and loss-of-function genomic screen-
ing. The chip could be tailored to different subgroups of the population and even to
individual patients by adjusting the expression levels of the various metabolizing
enzymes in human liver cells to match the enzyme levels representative of a subgroup
or unique to an individual. Using this technology, adverse responses of drug candi-
dates and their reactive metabolites by combinations of various DMEs in the human
liver can be assessed and quantified at speeds commensurate with early-stage human
toxicity tests. Ultimately, this approach would provide critical information needed
for the design of patient-specific treatment regimens, as well as for the identification
of pharmacologically safe and effective lead compounds for advancement to clinical
trials. With “high quality” information on chemical toxicity early on, pharmaceutical
companies can make better educated decisions on which compounds to take forward,
accelerate drug development times, and reduce investment in late-stage drug failures.
A list of microarray bioprinting-based assays is provided in Fig. 1.14, and more
information can be found in Chap. 8.
Combined
Chips
Enzyme, virus & Target
compound cells
References
1. Datar, A., Joshi, P., & Lee, M.-Y. (2015). Biocompatible hydrogels for microarray cell printing
and encapsulation. Biosensors, 5(4), 647–663. doi:10.3390/bios5040647.
2. Joshi, P., & Lee, M. Y. (2015). High content imaging (HCI) on miniaturized three-dimensional
(3D) cell cultures. Biosensors, 5(4), 768–790. doi:10.3390/bios5040768.
3. Lee, M.-Y., Park, C. B., Dordick, J. S., & Clark, D. S. (2005). Metabolizing enzyme toxicology
assay chip (MetaChip) for high-throughput microscale toxicity analyses. Proceedings of the
National Academy of Sciences, 102(4), 983–987. doi:10.1073/pnas.0406755102.
4. Lee, M. Y., & Dordick, J. S. (2006). High-throughput human metabolism and toxicity analysis.
Current Opinion in Biotechnology, 17(6), 619–627. doi:10.1016/j.copbio.2006.09.003.
5. Lee, M. Y., Clark, D. S., & Dordick, J. S. (2006). Human P450 microarrays for in vitro toxicity
analysis: Toward complete automation of human toxicology screening. Journal of the
Association for Laboratory Automation, 11(6), 374–380. doi:10.1016/j.jala.2006.08.003.
6. Lee, M.-Y., Dordick, J., & DS, C. (2010). Metabolic enzyme microarray coupled with minia-
turized cell-culture array technology for high-throughput toxicity screening. Methods in
Molecular Biology, 632, 221–237. doi:10.1007/978-1-60761-663-4.
7. Lee, D. W., Lee, M. Y., Ku, B., Yi, S. H., Ryu, J. H., Jeon, R., et al. (2014). Application of the
DataChip/MetaChip technology for the evaluation of ajoene toxicity in vitro. Archives of
Toxicology, 88(2), 283–290. doi:10.1007/s00204-013-1102-9.
8. Ziauddin, J., & Sabatini, D. M. (2001). Microarrays of cells expressing defined cDNAs.
Nature, 411(6833), 107–110. doi:10.1038/35075114.
9. Lee, M.-Y., Kumar, R. A., Sukumaran, S. M., Hogg, M. G., Clark, D. S., & Dordick, J. S. (2008).
Three-dimensional cellular microarray for high-throughput toxicology assays. Proceedings of
the National Academy of Sciences, 105(1), 59–63. doi:10.1073/pnas.0708756105.
10. Fernandes, T. G., Kwon, S. J., Bale, S. S., Lee, M. Y., Diogo, M. M., Clark, D. S., et al. (2010).
Three-dimensional cell culture microarray for high-throughput studies of stem cell fate.
Biotechnology and Bioengineering, 106(1), 106–118. doi:10.1002/bit.22661.
11. Fernandes, T. G., Kwon, S. J., Lee, M. Y., Clark, D. S., Cabral, J. M. S., & Dordick, J. S.
(2008). On-chip, cell-based microarray immunofluorescence assay for high-throughput analy-
sis of target proteins. Analytical Chemistry, 80(17), 6633–6639. doi:10.1021/ac800848j.
12. Kwon, S. J., Lee, D. W., Shah, D. A., Ku, B., Jeon, S. Y., Solanki, K., et al. (2014). High-
throughput and combinatorial gene expression on a chip for metabolism-induced toxicology
screening. Nature Communications, 5, 3739. doi:10.1038/ncomms4739.
13. Lee, D. W., Yi, S. H., Jeong, S. H., Ku, B., Kim, J., & Lee, M. Y. (2013). Plastic pillar inserts
for three-dimensional (3D) cell cultures in 96-well plates. Sensors and Actuators B: Chemical,
177, 78–85. doi:10.1016/j.snb.2012.10.129.
14. Lee, D. W., Choi, Y. S., Seo, Y. J., Lee, M. Y., Jeon, S. Y., Ku, B., et al. (2014). High-throughput
screening (HTS) of anticancer drug efficacy on a micropillar/microwell chip platform.
Analytical Chemistry, 86(1), 535–542. doi:10.1021/ac402546b.
15. Lee, D. W., Choi, Y.-S., Seo, Y. J., Lee, M. Y., Jeon, S. Y., Ku, B., et al. (2014). High-
throughput, miniaturized clonogenic analysis of a limiting dilution assay on a micropillar/
microwell chip with brain tumor cells. Small, 10(24), 5098–5105. doi:10.1002/smll.201401074.
16. Lee, D. W., Lee, M.-Y., Ku, B., & Nam, D.-H. (2015). Automatic 3D cell analysis in high-
throughput microarray using micropillar and microwell chips. Journal of Biomolecular
Screening, 20(9), 1178–1184. doi:10.1177/1087057115597635.
17. Kang, J., Lee, D. W., Hwang, H. J., Yeon, S.-E., Lee, M.-Y., & Kuh, H.-J. (2016). Mini-pillar
array for hydrogel-supported 3D culture and high-content histologic analysis of human tumor
spheroids. Lab on a Chip. doi:10.1039/C6LC00526H.
18. Sukumaran, S. M., Potsaid, B., Lee, M.-Y., Clark, D. S., & Dordick, J. S. (2009). Development
of a fluorescence-based, ultra high-throughput screening platform for nanoliter-scale cyto-
chrome p450 microarrays. Journal of Biomolecular Screening, 14(6), 668–678.
doi:10.1177/1087057109336592.
References 17
19. Kwon, S. J., Lee, M. Y., Ku, B., Sherman, D. H., & Dordick, J. S. (2007). High-throughput,
microarray-based synthesis of natural product analogues via in vitro metabolic pathway con-
struction. ACS Chemical Biology, 2(6), 419–425. doi:10.1021/cb700033s.
20. Zhang, H., Lee, M. Y., Hogg, M. G., Dordick, J. S., & Sharfstein, S. T. (2012). High-throughput
transfection of interfering RNA into a 3D cell-culture chip. Small, 8(13), 2091–2098.
doi:10.1002/smll.201102205.
21. Park, T.-J., Lee, M.-Y., Dordick, J. S., & Linhardt, R. J. (2008). Signal amplification of target
protein on heparin glycan microarray. Analytical Biochemistry, 383(1), 116–121. d oi:10.1016/j.
biotechadv.2011s.08.021.Secreted.
22. Kwon, S. J., Mora-Pale, M., Lee, M. Y., & Dordick, J. S. (2012). Expanding nature’s small
molecule diversity via in vitro biosynthetic pathway engineering. Current Opinion in Chemical
Biology, 16(1–2), 186–195. doi:10.1016/j.cbpa.2012.02.001.
Chapter 2
Microarray Spotter and Printing Technologies
Akshata Datar, Dong Woo Lee, Sang Youl Jeon, and Moo-Yeal Lee
Contents
2.1 Introduction........................................................................................................................ 20
2.1.1 Contact Printing Techniques.................................................................................. 20
2.1.2 Non-contact Printing Techniques........................................................................... 21
2.2 Materials............................................................................................................................ 22
2.3 Components of S+ MicroArrayer...................................................................................... 22
2.3.1 Main Body Components........................................................................................ 23
2.3.2 Externally Connected Components........................................................................ 26
2.4 General Precautions for Operating S+ MicroArrayer........................................................ 27
2.5 Daily Operations of S+ MicroArrayer............................................................................... 28
2.5.1 Turning ‘ON’ the System....................................................................................... 29
2.5.2 Refilling the Pressure Bottles................................................................................. 29
2.5.3 Washing Tubes, Solenoid Valves, and Ceramic
Tips with Ethanol and Water.................................................................................. 30
2.5.4 Dispensing Samples on the Micropillar/Microwell Chips
with a Work File........................................................................................... ......... 32
2.5.5 Replacing Solenoid Valves and Ceramic Tips....................................................... 34
2.5.6 Turning ‘OFF’ the System..................................................................................... 37
2.6 Detailed Programming for Normal Operation................................................................... 38
2.6.1 Generating Wash and Dry Sequences.................................................................... 38
2.6.2 Defining Well Plates............................................................................................... 38
2.6.3 Registering Chips................................................................................................... 41
2.6.4 Registering Spot Layouts....................................................................................... 44
2.6.5 Optimizing Dispensing Parameters Using Vision Inspection................................ 47
2.7 Summary............................................................................................................................ 50
References................................................................................................................................... 50
2.1 Introduction
The early stages of drug discovery heavily rely on high-throughput screening (HTS)
of compound libraries to identify effective lead compounds. Biochemical and
cell-based assays have been commonly performed with the assistance of robotic
liquid dispensing systems to rapidly test and identify the potential efficacy and tox-
icity of drug candidates [1, 2]. In an effort to save costs and reduce expensive
resources such as primary human cells and reagents, pharmaceutical industries have
focused on miniaturizing HTS assays by using higher density well plates including
384- and 1536-well plates, leading to a reduction in reagent volumes and an increase
in the speed of the liquid dispensing [1, 3].
Biological sample printing using microarray spotters is an important advance-
ment in the field of miniaturized assay development and HTS. Compared to tradi-
tional liquid dispensing systems, microarray spotters allow one to dispense
extremely small volumes (typically 200 pL to 950 nL) of biological samples (includ-
ing reagents, growth media, compounds, hydrogels, genes, proteins, viruses, and
cells) in microtiter plates, on glass slides, or on plastic chips. The process of cell
printing in hydrogels on a micropillar/microwell chip platform allows a precise
positioning of human cell spots, leading to the creation of more physiological rele-
vance of human cells grown in three dimensions (3D) [4, 5].
In this chapter, we will briefly introduce the advantages of various printing tech-
nologies and go over general precautions that have to be taken when printing various
biological samples with a microarray spotter. In general, microarray spotting is
divided into direct contact printing using pins and non-contact printing using solenoid
valves, piezoelectric nozzles, etc. [6, 7]. Although we have experience in operating
several microarray spotters, including MicroSys and PixSys from DigiLab and Nano-
Plotter from GeSim, we will provide detailed protocols on how to operate S+
MicroArrayer from Samsung Electro-Mechanics, Co. (SEMCO). S+ MicroArrayer
represents the most advanced microarray spotter for cell printing, which is specifically
designed to accommodate the micropillar/microwell chip. The principle of operating
and troubleshooting S+ MicroArrayer can be applicable to any solenoid-driven micro-
array spotters.
the tip comes in direct contact with a liquid. Several types of tips have been
developed, depending on the material used and the capillary size inside the tip to
manage the volume of the droplet printed [8]. Factors that determine the volume of
the droplet are the capillary size, surface tension of the liquid, affinity of the liquid
with the glass slide, and the surface chemistry of the glass slide. Pin-based printing
is mainly used to print extremely small volumes of proteins and DNA on the sur-
face of glass slides [9, 10]. Although pin-based printing is straightforward and fast,
it is difficult to print colloidal suspensions (such as cells), and the droplet size may
be inaccurate and inconsistent depending on surface and liquid properties [9]. For
dip-pen nanolithography, AFM is used to deposit some biological samples, mostly
small molecules on the surface [7, 11]. Although this technology has capable of
printing small spots in 25–200 nm resolution, it is not widely used in printing
biological samples due to lack of multiplexing capabilities and limited detection
methods [7, 9].
In non-contact printing, microarray spotters with solenoid valves are the most com-
monly used, which function on the principle of electromagnetic induction. Typical
solenoid valves consist of a metal rod that is surrounded by coiled wires that con-
duct an electric current. When voltage is applied across the coils, a magnetic field is
generated that forces the metal rod to act as a shaft that moves up and down. This
metal rod acts as a gate while printing biological samples. The setup is enclosed in
a nozzle that has constant sample supply and constant pressure maintained using
syringe pumps. When the shaft moves up and down, the outlet for printing biologi-
cal samples opens and closes. This is how the voltage applied to the solenoid valves
controls dispensing of liquid samples [6]. Unlike other dispensing systems, the sole-
noid valves can handle colloidal samples quite well. Typical dispensing volumes are
30–950 nL per droplet. One particular printer, S+ MicroArrayer, is equipped with
six solenoid valves for printing six samples simultaneously.
Piezoelectric nozzles that are used in conventional inkjet printers are also
commonly used in microarray printing with dispensing volume of 200–600 pL per
droplet. Unlike solenoid valves, the dispensing volume is controlled by the inner
structure of piezoelectric nozzles. Microarray spotters with piezoelectric nozzles
use electric pulses that control the expansion and contraction of a piezoelectric
membrane that acts as a pump to push biological samples through the tip [12, 13].
Thus, the dispensing volume can be increased not by changing voltages, but by
depositing multiple spots at the same location. Although piezoelectric nozzles are
capable of printing biological samples, its use is limited due to inconsistent printing
of colloidal samples and difficulty of controlling large dispensing volumes [14].
Piezoelectric nozzles are more commonly used for printing compounds dissolved in
DMSO, proteins, and amino/nucleic acids.
22 2 Microarray Spotter and Printing Technologies
2.2 Materials
To better explain the operation of the microarray spotter, we will classify it under
mechanical and software components. Although both components work hand-in-
hand, their maintenance and parameter setting are different from each other.
Mechanical components are divided into the main body, the utility body, and exter-
nally connected parts that play an important role in chilling, rinsing, and sample
dispensing (Figs. 2.1 and 2.2).
2.3 Components of S+ MicroArrayer 23
Main body
Pressure bottle
Utility body
Waste bottle
Chillers
The main body components are housed in a chamber where biological samples are
loaded in a 96-well plate and printed on a micropillar/microwell chip at controlled
temperature and humidity. It mainly consists of a robotic arm, a dispensing head with
six solenoid values and ceramic tips, six syringe pumps, a droplet inspection camera,
a chip inspection camera, two chip-loading decks, a well plate deck, a vacuum pump,
a humidifier, a water bath with a sonicator, and a waste drainage basin (Fig. 2.3).
• Dispensing head unit
The dispensing head unit of the microarray spotter is mainly responsible
for moving and printing biological samples in accurate volumes and positions.
It consists of a chip alignment inspection camera, six solenoid valves connected
to syringe pumps through tubing, and six ceramic tips (Fig. 2.4). It is capable of
moving X, Y and Z directions, hence aspirating biological samples from
96-well/384-well plates and dispensing in different patterns on the micropillar/
microwell chip.
–– Solenoid valves: They are the main component behind microarray bioprinting
technology, which function on the principle of electromagnetic induction.
The voltage applied to the valves, creates a magnetic field that forces the gate
to open and close. The syringe pumps maintaining the pressure help the
samples to flow when the gate is open. Each solenoid valve can be controlled
individually.
24 2 Microarray Spotter and Printing Technologies
Controller
Computer
Humidity indicator
Air pressure
gauge
Reset button
Emergency
button
Syringe pump
Z
axis Sonicator
Chip deck
Solenoid valve
Sonicator
Chip alignment
inspection camera
Ceramic tips
Plate deck
tained by the syringe pumps forces the samples out on the micropillar and
microwell chips.
• Droplet inspection camera: There are two cameras in the main body of the
microarray spotter. The chip alignment inspection camera is installed in the
dispensing head, and the droplet inspection camera is placed beside the chip-
loading deck, which is used to optimize dispensing parameters such as air
26 2 Microarray Spotter and Printing Technologies
pressure and solenoid valve open time. When sample droplets are dispensed on
a hydrophobic plastic strip, the droplet inspection camera takes pictures of the
droplets and calculate dispensing volumes at the setting by measuring the height
and diameter of the droplet. The dispensing setting and actual dispensing volume
can be compared for parameter optimization.
• Water bath with a sonicator and a vacuum pump: These two components are
sequentially used to clean the tips and tubing that hold the samples between
aspiration and dispensing of samples. The water bath with a sonicator is respon-
sible for rinsing the surface of the tips and breaking down any bigger particles
that tend to settle and clog the tips using ultrasound. The vacuum pump is respon-
sible for drying the tips to prevent sample carry-over.
• Pressure bottles: One pressure bottle contains water for sample printing and the
other bottle holds 70 % ethanol for rinsing and sterilization (Fig. 2.5). The tubing
has to be filled with water to transfer energy from the syringe pumps and push
the samples to flow for printing. The bottles are pressurized with air to supply
water and alcohol.
• Chillers: There are two chillers equipped in the microarray spotter, one for the
dispensing head for printing temperature-sensitive samples and the other for the
chip-loading decks for preventing spot drying, the 96-well plate deck, and the
two pressure bottles (Fig. 2.6).
• Waste bottle: It holds wastewater generated by rinsing and washing the solenoid
valves, the ceramic tips, and the tubing.
Water outlet
Ethanol outlet
2.4 General Precautions for Operating S+ MicroArrayer 27
• Solenoid valves can be used to print viscous solutions, but may not be suitable
for spotting organic solvents except alcohols because of incompatible plastic
parts in the solenoid valve. Aspirating DMSO with solenoid valves for a short
period of time will be okay. When compounds in DMSO is diluted with solenoid
valves, use a 384-well plate, and do not aspirate more than 20 μL of samples.
• Do not attempt to print spontaneously gelling materials in any case. In case
of printing cells in Matrigel, always maintain the dispensing head at low tem-
perature (4 °C) to prevent the gelation of Matrigel inside of the solenoid valves
while printing. An extensive rinse of tubes and solenoid valves with ice-cold
water is critical before and after printing Matrigel solution.
• Keep in mind that there is no “space” allowed when saving a file name, and do
not use special characters as well.
• If you have a new operational file made, test it without ceramic tips installed first.
Do not install the expensive tips for testing unknown work files.
Chip window
Display
window
IO window
Chip alignment
inspection window Droplet inspection
window
Command window
Fig. 2.7 The main screen of ezAOI. The Command window has four groups—Program,
Equipment setup, User operation, and Work
2.5 Daily Operations of S+ MicroArrayer 29
samples on the micropillar and microwell chips in distinctive patterns. The main
screen that pops up displays six windows, including command window, slide
window, IO window, display window, chip alignment inspection window, and
droplet inspection window. The command window in itself has four tabs, includ-
ing program, equipment setup, user operation, and work. Each of these tabs has
different functions, which allow us to program the microarray spotter more
efficiently. For daily operation of the microarray spotter, follow the protocols
provided below.
1. Open the air cylinder (or in-house air valve) and maintain the pressure of com-
pressed air at 100 psi.
2. Turn on the chiller. Note: For most of applications, there is no need to turn on
the humidifier because of surface cooling. Maintain the temperature of the
chip-loading deck between 4–10 °C to reduce evaporation of water in spots on
the chip. It is extremely important to avoid excess water condensation on the
bottom layer to prevent spot detachment.
3. Turn on the external switch in the utility body.
4. Turn on the computer and the monitor.
5. Prior to running the ‘ezAOI’, push ‘Reset button’ and reset XYZ coordinates.
Note: This step is essential to avoid malfunctioning of the microarray spotter.
Do not skip this step!
6. Run the ‘ezAOI’ software. The software will initialize the system automatically
when running. Note: Make sure that no obstacles are on the work place to
avoid the robotic arm crashing.
6. Close the lid of the pressure bottles carefully. Note: Make sure that the ‘O’ ring
is properly placed before placing the lid.
7. Turn the orange knob of the pressure bottles from ‘O’ to ‘S’ position, which
close the pressure bottles completely.
8. Click ‘Initialize’ of the ‘Air pressure’ box in the ‘Water & Alcohol Filling Up’
box to apply air pressure in the bottles.
and solenoid valves with ice-cold water is critical before and after printing
Matrigel solutions.
6. Check a streamline of water while washing. Note: If the streamline is deflected
from the vertical or water is beaded up at the tip end, clean the ceramic tip with
sonication.
32 2 Microarray Spotter and Printing Technologies
Fig. 2.11 The images of the original micropillar shown as “white spot” (left) and the micropillar
with the BaCl2-PLL bottom layer shown as “black spot” (right)
3. Select the well plate used for sample loading and aspiration in ‘Well Plate ID’.
Note: It is extremely important to use the same kind of the 96-well plates
always to avoid tip crashing due to Z height difference in 96-wells. Enter the
offset values in the X and Y directions to designate the location of samples in
96-wells. The exact location of sample wells will be appeared in the schematic
of the well plate in the ‘Display’ window.
4. Select the type of a chip in ‘Slide/Chip ID’ onto which samples are dispensed.
The schematic of the chip selected will be appeared in the ‘Chip’ window.
5. Select the color of spots on the chip under the camera in ‘Spot Color’. Typically,
it is white when nothing is dispensed on the chip or black when BaCl2/PLL is
dispensed (Fig. 2.11). Note: This color selection is necessary to automatically
identify the location of the chip with the chip alignment inspection camera.
6. Enter the air pressure used for sample spotting and the desired droplet volume
for solenoid valves in ‘Spot Volume’. Note: The typical air pressure used is 6
kilopascal (kPa) for 40–950 nL droplets.
7. Enter the air gap between the sample and water. Note: The typical air gap used
is 0 μL for 40–150 nL droplets and 20 μL for 700–950 nL droplets. For accu-
rate sample spotting, small or no air gap is allowed for small dispensing
volumes.
8. Select the wash and dry sequences you want before and after sample dispensing
in ‘Wash Dry’.
9. Select ‘No inspection’ in ‘Camera Inspection ID’. Note: For rapid sample dis-
pensing, we typically skip camera inspection of droplets. Camera inspection
of sample spotting is necessary to optimize dispensing parameters at first.
10. Enter the speed of the axis movement for solenoid valves in ‘Speed of Axis
Movement’. It is typically 10–20 mm/s.
11. Open the optimum dispensing parameters of solenoid valves obtained from the
vision inspection in ‘Optimum Tip Parameter’. Note: Manually change the
open time, if needed. Optimum dispensing parameters will not be changed for
each sample and droplet volume used, unless the solenoid valves are broken
34 2 Microarray Spotter and Printing Technologies
In case of ceramic tips clogging and solenoid valves malfunctioning, ceramic tips
and solenoid valves have to be removed and replaced. Prior to removing tips and
solenoid valves, run ‘Solenoid Fix’, force to dispense large droplets multiple times
(typically 500 times), and rinse solenoid valves with water and sonication.
1. Select the ‘User Operation’ window.
2. Select ‘Wash & Dry’ under Daily Operation options (Fig. 2.12).
3. Select ‘Solenoid Fix’ and select all the nozzles from 1 through 6.
2.5 Daily Operations of S+ MicroArrayer 35
Fig. 2.14 The pictures of the dispensing head for replacing ceramic tips and solenoid valves
15. Attach a new ceramic tip to the solenoid valve, place them accordingly, and
connect the wire back in.
16. Before putting the screws in, make sure that the ceramic tips are leveled on the
metal block.
17. If the new ceramic tip still doesn’t print samples properly, the solenoid valve is
clogged, thus replacing it with a new one. Note: Run ‘Daily Washing’ after
replacing solenoid valves and before printing samples.
1. To avoid potential contamination issues, rinse tubes and solenoid valves with
ethanol and water thoroughly by running ‘Daily Washing’.
2. Close ‘ezAOI’ software.
3. Turn off the computer and the monitor.
4. Turn off the external power switch in the utility body.
5. Turn off the chillers. Note: Make sure to turn off the chillers to avoid excess
water condensation on the chip-loading deck and the dispensing head, causing
a short circuit by water.
6. Close the air cylinder (or in-house air valve). Note: Do not close the air cylinder
while ‘ezAOI’ is on.
38 2 Microarray Spotter and Printing Technologies
These sections are prepared for advanced users who want to know how to program
the operation of S+ MicroArrayer.
Note: This will be the position of the first ceramic tip for aspirating samples.
Do not change Z levels randomly to prevent the tips crashed into the well
surface.
8. Select ‘Well Plate’ in the ‘Program’ box again.
9. Enter ‘Well A1 Position’ determined from Step 7.
10. Enter minimum and maximum sample volumes allowed in either 96 or 384 wells.
11. Click ‘Save as’ button to register well plate information.
12. Enter a file name. Note: There is no space and special character allowed in
the file name.
40 2 Microarray Spotter and Printing Technologies
13. If you want to modify the well plate you registered, select the file name in the
box above the ‘Save as’ button and repeat Steps 3, 4, 9, and 10.
14. Click the ‘Save’ button to overwrite the changes in the same file.
5. Click ‘1st Spotting Layout’ button to generate the first spotting pattern on the
chip with samples.
6. Select ‘Solenoid’ in the ‘Select Nozzle’ box as by default we use it for sample
spotting (Fig. 2.23).
7. Select the number of solenoid valves/ceramic tips and the layout used in the
‘Nozzle Layout’ box.
8. Select the type of the well plate used (96 or 384).
9. Select a 96-well (or a 384-well) where the first solenoid valve/ceramic tip will
be located for sample aspiration by clicking the well. According to the number/
layout of solenoid valves used, other wells will be automatically selected.
46 2 Microarray Spotter and Printing Technologies
Note: For safety reasons, the entire well will not be selected when any one of
the nozzles is left outside of the well plate for aspiration.
10. Select the region of spots where the sample will be dispensed with the first
solenoid valve/ceramic tip by clicking and dragging on the chip layout.
11. Repeat Steps 9 and 10 according to the spot layout you want to generate.
12. Select the box beside ‘Set Sacrificial Spot’ to designate a sample well for print-
ing sacrificial regions. Note: Always the first solenoid valve/ceramic tip is used
for spotting sacrificial regions.
13. Select the sample well for spotting sacrificial regions.
14. Click ‘Close’ button when done.
15. Click ‘Save as’ button to register the spot layout.
2.6 Detailed Programming for Normal Operation 47
16. Enter a file name. Note: There is no space and special character allowed in
the file name.
17. If you want to modify the spot layout you registered, click the ‘Open’ button
and select the file name and repeat Steps 3 through 14.
18. Click the ‘Save’ button to overwrite the changes in the same file.
Fig. 2.23 The screen of ‘Program’ > ‘Spot Layout’> ‘1st Spotting Layout’
7. Select tip rinsing conditions in water before and after sample loading prior to
sample dispensing in the ‘Before’ and ‘After’ boxes.
8. Enter the droplet volume for inspection in ‘Dispensing Volume’, the open time
of the solenoid valves selected in ‘Open Time’, and the number of spots dis-
pensed to test the conditions in ‘Number of Spot’ in the ‘Manual Dispensing
Setup’ box. Note: If the disparity between the droplet volume set up and the
average volume of the measured droplets at a certain condition is less
than the desired CV value (typically 5 %), then the vision inspection will be
successfully finished.
50 2 Microarray Spotter and Printing Technologies
9. After loading test samples, click the ‘Dispense’ button to test the dispensing
parameters manually with the camera.
10. The open time of each solenoid valve will be updated when the vision inspection
is successfully completed; otherwise 0 will be reported in the ‘Open Time’ box.
Repeat Steps 3–9 when failed.
11. After finishing all solenoid valve inspection, click ‘Save as’ button to save the
optimum dispensing parameters.
12. Enter a file name. Note: There is no space and special character allowed in
the file name.
13. If you want to modify the optimum dispensing parameters you saved, click the
‘Open’ button, select the file name, and then repeat Steps 3 through 11.
14. Click the ‘Save’ button to overwrite the changes in the same file.
2.7 Summary
References
1. Dunn, D. A., & Feygin, I. (2000). Challenges and solutions to ultra-high-throughput screening
assay miniaturization: Submicroliter fluid handling. Drug Discovery Today, 5(12), 84–91.
doi:10.1016/S1359-6446(00)80089-6.
2. Kapur, R., Giuliano, K. A., Campana, M., Adams, T., Olson, K., Jung., D., et al. (1999).
Streamlining the drug discovery process by integrating miniaturization, high throughput
screening, high content screening, and automation on the CellChip TM system. Biomedical
Microdevices, 2(2), 99–109.
3. Horman, S. R., To, J., Orth, A. P., & Caracino, D. (2013). High-content analysis of three-
dimensional tumor spheroids: Investigating signaling pathways using small hairpin RNA.
Nature Methods, 10(10), v–vi. doi:10.1038/nmeth.f.370.
4. Justice, B. A., Badr, N. A., & Felder, R. A. (2009). 3D cell culture opens new dimensions in
cell-based assays. Drug Discovery Today, 14(1), 102–107. doi:10.1016/j.drudis.2008.11.006.
5. Comley, J. (2013). Progress made in applying 3D cell culture technologies. Drug Discovery
World., 15, 41–58.
References 51
6. Datar, A., Joshi, P., & Lee, M.-Y. (2015). Biocompatible hydrogels for microarray cell printing
and encapsulation. Biosensors, 5(4), 647–663. doi:10.3390/bios5040647.
7. Salaita, K., Wang, Y., & Mirkin, C. A. (2007). Applications of dip-pen nanolithography.
Nature Nanotechnology, 2(3), 145–155. doi:10.1038/nnano.2007.39.
8. Cho, E. J., & Bright, F. V. (2002). Integrated chemical sensor array platform based on a light
emitting diode, xerogel-derived sensor elements, and high-speed pin printing. Analytica
Chimica Acta, 470(1), 101–110. doi:10.1016/S0003-2670(02)00303-3.
9. Mota, C., & Moroni, L. (2015). Chapter 11—High throughput screening with biofabrication
platforms. In A. Atala & J. J. Yoo (Eds.), Essentials of 3D biofabrication and translation
(pp. 187–213). Amsterdam: Elsevier. doi:10.1016/B978-0-12-800972-7.00011-6.
10. Dusseiller, M. R., Schlaepfer, D., Koch, M., Kroschewski, R., & Textor, M. (2005). An inverted
microcontact printing method on topographically structured polystyrene chips for arrayed
micro-3-D culturing of single cells. Biomaterials, 26(29), 5917–5925. doi:10.1016/j.
biomaterials.2005.02.032.
11. Zhang, L. G., Fisher, J. P., & Leong, K. W. (2015). 3D Bioprinting and Nanotechnology in
Tissue Engineering and Regenerative Medicine. Amsterdam: Elsevier. doi:10.1016/
B978-0-12-800547-7.00004-7.
12. Boland, T., Xu, T., Damon, B., & Cui, X. (2006). Application of inkjet printing to tissue engi-
neering. The Biochemical Journal, 1(9), 910–917. doi:10.1002/biot.200600081.
13. Cui, X., Boland, T., D’Lima, D. D., & Lotz, M. K. (2012). Thermal inkjet printing in tissue
engineering and regenerative medicine. Recent Patents on Drug Delivery & Formulation, 6(2),
1–13. doi:10.2174/187221112800672949.
14. Nishiyama, Y., Nakamura, M., Henmi, C., Yamaguchi, K., Mochizuki, S., Nakagawa, H., et al.
(2009). Development of a three-dimensional bioprinter: construction of cell supporting struc-
tures using hydrogel and state-of-the-art inkjet technology. Journal of Biomechanical
Engineering, 131(3), 1–6. doi:10.1115/1.3002759.
15. PK, W., Ringeisen, B. R., Callahan, J., Brooks, M., Bubb, D. M., Wu, H. D., et al. (2001).
The deposition, structure, pattern deposition, and activity of biomaterial thin-films by matrix-
assisted pulsed-laser evaporation (MAPLE) and MAPLE direct write. Thin Solid Films, 398–399,
607–614. doi:10.1016/S0040-6090(01)01347-5.
16. Devillard, R., Pagès, E., Correa, M. M., Kériquel, V., Rémy, M., Kalisky, J., et al. (2014). Cell
patterning by laser-assisted bioprinting. Methods in Cell Biology, 119, 159–174. doi:10.1016/
B978-0-12-416742-1.00009-3.
17. Catros, S., Guillotin, B., Bačáková, M., Fricain, J. C., & Guillemot, F. (2011). Effect of laser
energy, substrate film thickness and bioink viscosity on viability of endothelial cells printed by
Laser-Assisted Bioprinting. Applied Surface Science, 257(12), 5142–5147. doi:10.1016/j.
apsusc.2010.11.049.
18. Demirci, U., & Montesano, G. (2007). Single cell epitaxy by acoustic picolitre droplets.
Lab on a Chip, 7(9), 1139–1145. doi:10.1039/b704965j.
19. Fang, Y., Frampton, J. P., Raghavan, S., Sabahi-Kaviani, R., Luker, G., Deng, C. X., et al.
(2012). Rapid generation of multiplexed cell cocultures using acoustic droplet ejection fol-
lowed by aqueous two-phase exclusion patterning. Tissue Engineering. Part C, Methods,
18(9), 647–657. doi:10.1089/ten.tec.2011.0709.
20. Demirci, U., & Montesano, G. (2007). Cell encapsulating droplet vitrification. Lab on a Chip,
7(11), 1428–1433. doi:10.1039/b705809h.
Chapter 3
Chip Platforms and Chip Surface Treatments
Contents
3.1 Introduction........................................................................................................................ 54
3.1.1 Surface Modification of Micropillar and Microwell Chips................................... 54
3.1.2 Surface Modification of Glass Slides..................................................................... 56
3.2 Materials............................................................................................................................ 57
3.3 Protocols............................................................................................................................ 58
3.3.1 Cleaning Microwell Chips with Plasma for Air Bubble Removal......................... 58
3.3.2 Coating Micropillar Chips with Poly(maleic anhydride-alt-1-octadecene)
(PMA-OD) for Cell Printing.................................................................................. 61
3.3.3 Cleaning the Surface of Glass Slides with Acid.................................................... 61
3.3.4 Coating of Glass Slides with Poly(styrene-co-maleic anhydride)
(PS-MA) for Cell Printing..................................................................................... 63
3.3.5 Coating of Acid-Cleaned Glass Slides
with Methyltrimethoxysilane (MTMOS) for Enzyme Printing............................. 64
3.3.6 Coating of Acid-Cleaned Glass Slides with (3-Aminopropyl)trimethoxysilane
(APTMS) for Protein Attachment......................................................................... 65
3.3.7 Measurement of Silanization on the Surface of the APTMS Slides
with Fluorescein Isothiocyanate (FITC) Labeling................................................. 65
3.3.8 Attachment of Glutaraldehyde on the Surface of the APTMS Slides
for Protein Attachment.......................................................................................... 66
3.3.9 Immobilization of Extracellular Matrix (ECM) Proteins
on Reactive Surfaces.............................................................................................. 67
3.4 Summary............................................................................................................................ 68
References................................................................................................................................... 69
3.1 Introduction
There are three significant types of platforms used in microarray printing technol-
ogy: glass slides, micropillar chips, and microwell chips. The surfaces of these plat-
forms are treated for different applications. For example, a glass slide is ideal for
micropatterning and can be used for microscopic image analysis. Micropillar and
microwell chips are ideally suited for applications of high-throughput screening
(HTS) of compounds on 3D-cultured cells. In this chapter, we present various
methods by which these surfaces can be chemically modified for use in bioprinting
applications, with emphasis on the methods used for culture of cells encapsulated
in hydrogels.
Since many cells have difficulty in attaching to the surface of glass slides, different
coatings on glass slides have been developed to improve cell attachment and growth,
such as using rat tail collagen. In addition, glass slides need very tedious and precise
cleaning procedures to make sure that any residues on the surface do not affect cel-
lular processes. To alleviate these issues and facilitate robust and reproducible cell
cultures, our research team developed disposable plastic chip platforms such as a
micropillar chip and a microwell chip (Fig. 3.1).
Both micropillar and microwell chips are made of polystyrene, a long carbon
chain polymer with benzene rings attached to every other carbon (Fig. 3.2).
Polystyrene has a very good optical clarity, is easily moldable, and can be sterilized
by UV or plasma irradiation. However, it is hydrophobic in nature without func-
tional groups; thus, cells and hydrogels with cells have difficulty in attaching on its
surface. In order to have proper attachment, the surface of polystyrene is modified
Fig. 3.1 Schematics of (A) a micropillar chip and (B) a microwell chip. (C) Image of micropillar
and microwell chips to scale against a microscopic glass slide
3.1 Introduction 55
PLL
PMA-OD
Polystyrene
Fig. 3.3 Modification of polystyrene surface via plasma treatment. A variety of different chemical
groups can be introduced by breaking the carbon chain backbone
strong gas plasma under vacuum. This process generates highly energetic oxygen
ions and radicals, which oxidizes the polystyrene chains (Fig. 3.3) so that the
surface becomes hydrophilic and negatively charged [2–5]. This process uses an
ionized gas or plasma, to remove all organic matter from the surface of an object,
which is carried out in a vacuum chamber. During plasma treatment, excited gas
molecules and atoms in the plasma will emit UV light. By introducing different
gases into the plasma chamber such as O2, N2, Ar, He, and CF4, treatment can be
performed. There are no harsh chemicals involved in the process of plasma cleaning;
therefore, it is an environmentally safe process.
Glass slides have significant use for biological testing. The distinct advantages of
glass substrates are that they are relatively non-reactive, which is suitable for bio-
compatibility, and they are well suited for imaging purposes. Additionally, glass
slides are thermally stable, which makes them relatively easy to clean and modify
for future uses. The main struggle with glass surfaces is that their non-reactivity,
while potentially useful for culturing cell monolayers, makes them a poor surface
for hydrogel attachment. This hydrogel attachment is critical, as hydrogels and
other structures have the potential for cells to organize themselves into three-
dimensional (3D) structures, mimicking the cell-ECM interactions found in vivo.
If hydrogels are not properly attached, any treatment towards the glass slide can
result in subsequent loss of cells encapsulated in hydrogels. Thus, it is a goal of
researchers to investigate methods for robustly immobilizing hydrogels onto glass
slides without compromising the integrity of the hydrogel or the normal behavior
of the cells.
Silicone precursor molecules are important to initiate this attachment between
glass slides and biological molecules. The strong covalent bonding between the
glass made of silicone oxide and the silicone precursor allows for the interaction of
biologics (such as hydrogels, proteins, and cell surfaces) with the functionalized
glass slide with the silicone precursor. One potential method that has been used to
improve protein and hydrogel attachment onto glass slides is the use of silanes with
amine groups, which can covalently attach biological compounds to silane surfaces.
The silicon in the silane group anchors well with the silicon glass slide, while func-
tional groups attached to the silane can interact with functional groups on hydrogels
and cell membranes to promote robust attachment. Such functional groups include
thiols, esters, N-hydroxysuccinimide (NHS)-esters, amines, and methacrylates [6].
Seo et al. demonstrated the use of thiol silanes and acryl silanes for the covalent
attachment of polyethylene glycol (PEG) hydrogels and streptavidin onto glass
slides without compromising the hydrogel integrity or the affinity of the streptavidin-
biotin interaction [6]. While silanes with initial functional groups can be maintained
on a glass surface, H-terminal silanes can also be modified. Carvalho et al. demon-
strated the conjugation of alkenes with H-terminal silanes on glass via ultraviolet
(UV) light [7]. Luo et al. used a similar procedure, but applied acrylamide onto the
3.2 Materials 57
surface for a different surface chemistry [8]. The use of UV light is important for
treating glass surfaces, as it serves the dual function of modification of the surface,
and sterilization.
While UV light has been used as a mechanism for chemical modification of glass
surfaces, the thermal stability of silicon makes glass a suitable surface for thermally
controlled surface modification. Cheawchan et al. has demonstrated this by initially
modifying the glass surface with allyl groups before 70 °C heating was applied, and
the surface was reacted with nitrile n-oxide moieties [9]. The glass slide can then be
further modified for addition of functional groups that can aid in the attachment of
cells and hydrogels to the glass surface.
Another potential method for improving the attachment of biological molecules
onto glass slides involves initial surface treatments with gold. Gold can be easily
coated onto the surface of glass slides using various physical methods, but vapor
deposition is among the most common ways for entirely plating gold onto a glass
slide [10]. This is followed up with subsequent attachment of thiol groups [10].
The covalent attachment of sulfur on gold is easily achievable. This means that cysteine
and potentially methionine residues can be targeted on proteins for immobilization
purposes. Additionally, disulfide linkages can be used for a potential point of interac-
tion between gold and any thiol residues on the surface of the glass slide.
3.2 Materials
3.3 Protocols
Plasma treatment of microwell chips is carried out to change the surface property of
the microwell chip hydrophilic and remove air bubbles trapped in the microwells
when immersed in growth media or to sterilize the microwell chips. The protocol of
plasma treatment provided here is modified from the user manual from the manu-
facturer [11].
1. Make sure the machine and the thermocouple are plugged into an appropriate
power source, and make sure the vacuum pump is plugged into the plasma
cleaner. Turn on the power switch located on the vacuum pump, but do not touch
the ‘PUMP’ switch on the Plasma Cleaner’s panel just yet.
2. Turn on the Plasma Cleaner’s main power (‘POWER’ switch to ON).
3. Open the door of the Plasma Cleaner. Load microwell chips on the Pyrex glass
tray and put it inside the Plasma Cleaner’s chamber. Maximum ten chips can
fit on the tray (Fig. 3.4). Note: Place the chips facing up in the position for
plasma treatment. In addition, wipe the back of the microwell chips with
70 % ethanol-soaked paper towels to prevent microbial contamination.
4. Check that the 3-way valve is closed. It must be leveled in the vertical downward
position (Fig. 3.5).
5. Close and lean the door tightly against the Plasma Cleaner.
6. While holding the door, set the Plasma Cleaner’s ‘PUMP’ switch to ON. It is
located directly on the Plasma Cleaner (Fig. 3.6). The vacuum will maintain the
door tightly closed. Wait for 5 min.
7. Level the 3-way valve into the horizontal position such that it points to the left
(Fig. 3.7).
8. Slightly open the metering valve and wait for 1 min (Fig. 3.7). It will allow the air
to enter the Plasma Cleaner. Note: The metering valve should be operated in the
way the plasma intensity is noticeably increased and the pressure at the thermo-
couple is ranging from 800 to 1000 mTorr. Other pressures may also work
depending on the samples, but 900–1000 mTorr is ideal for microwell chips.
9. Turn on Plasma Cleaner’s RF switch to ‘HI’. Wait for 1 min. A fluorescence should
be observed through the Plasma Cleaner’s window. For a plasma cleaning process
with room air, purple-pink to violet plasma should be observed (Fig. 3.8).
Fig. 3.6 Plasma treatment of microwell chips is performed using (A) a vacuum pump, (B) a digi-
tal thermocouple gauge control unit, and (C) a plasma cleaner unit
60 3 Chip Platforms and Chip Surface Treatments
10. Process samples for 15 min. Note: The process time has to be optimized for
different applications.
11. At the end of the process, switch the RF level to ‘OFF’.
12. Level the 3-way valve to the vertical downward position (Fig. 3.5)
13. Turn off the vacuum pump by setting the Plasma Cleaner’s ‘PUMP’ switch
to OFF.
14. Slowly level the 3-way valve into the horizontal position such as it points to the
right (Fig. 3.9). The system will vent. Note: The Plasma Cleaner must be vented
3.3 Protocols 61
instantly. Otherwise, oil from the vacuum pump may back stream from the
pump and contaminate the system. Do not open the front door when the chamber
is under vacuum since this will damage the glass sample holder and any samples.
Additionally, a slow vent ensures that the samples and the Pyrex tray do not move
around the chamber when the samples are removed.
15. When the hissing noise from venting the chamber is not observed anymore,
open the front door.
16. Close the 3-way valve. Level it to the vertical downward position (Fig. 3.5).
17. Turn off the Plasma Cleaner’s main power (‘POWER’ switch to OFF).
18. Remove the chip samples.
19. Unplug the thermocouple at the end of the process, and keep it unplugged when
it is not used.
For a very even silanization of a glass surface as well as uniform polymer coating,
removing all dirt and exposing silanol group (─SiOH) on the surface of glass slides
with strong acid are extremely important [12]. Unless pre-cleaned precision glass
62 3 Chip Platforms and Chip Surface Treatments
Fig. 3.10 Coating the surface of a micropillar chip with 0.01 % (w/v) PMA-OD in a shallow
staining plate
slides from Coresix Precision Glass, Inc. are used, the acid-cleaning step is essential
for standard microscope slides.
1. Properly number plain microscope glass slides (25 × 75 × 1 mm) with a
diamond tip.
2. Wipe the glass slides three times with 70 % ethanol-soaked paper towels to
remove oil and solid particles on the slide surface and then clean ethanol with
3.3 Protocols 63
dry paper towels. Note: Do not use Kimwipes because it leaves lots of small
paper particles. After wiping with dry paper towels, the microscope slides
should not be wet with ethanol. It will leave ethanol stains on the slide surface.
3. Place the glass slides in a removable glass rack, immerse the rack in a Wheaton
glass dish filled with concentrated sulfuric acid (96.5 %) for 2 h, and then soni-
cate the Wheaton glass dish with the slides in acid for 10 min. Note: Always
work in a chemical fume hood while handling acid or organic solvent, do not
inhale acidic fume, and do not reuse sulfuric acid more than twice. Dispose of
all chemicals in a labeled waste container properly.
4. Rinse the rack containing the glass slides at least five times with de-ionized dis-
tilled water to remove excess sulfuric acid on the slide surface.
5. Immerse the rack containing the slides twice in a Wheaton glass dish filled with
acetone to remove excess water on the glass slide and dry the acid-cleaned glass
slides thoroughly under N2 gas stream.
6. Bake the acid-cleaned slides in an oven at 120 °C for 10 min to completely
remove water on the slide surface.
7. After drying, carefully inspect the glass slide for cleanliness. If the slide is not
clean enough (i.e., any dirt or particle on the slide surface), return to the Step 2.
Note: Always use tweezers after the Step 2. Do not use fingers to hold the
slides, as this will leave an oily residue on the slide.
8. Store the glass slides in a sterile petri dish (150 mm diameter, maximum 5 slides/
petri dish) or the bioassay plate until use. Never leave the clean slide uncovered
for long time.
To attach cell spots or prepare cell patterns on the surface of glass slides, it is
required to modify the surface amine-reactive while enhancing hydrophobicity to
prevent spot spreading. The amphiphilic PS-MA coating is suitable to change the
hydrophilic surface of glass slides to hydrophobic while providing amine-reactive
functionality for cell spot attachment.
1. Prepare a fresh 1 % (w/v) solution of PS-MA by dissolving PS-MA in anhydrous
toluene in a scintillation vial with a PTFE-coated lid. Note: Use disposable glass
serological pipets to aspirate toluene precisely.
2. Shake the solution in an incubating shaker at 50 °C, 150 rpm for 40 min until
PS-MA pellets completely dissolved. Note: After 30 min of shaking at 50 °C,
150 rpm, the solution usually become clear. However, leave the solution for
additional 10 min for complete dissolution of PS-MA pellets. The PS-MA solu-
tion should be prepared freshly as reactivity of maleic anhydride groups
decreases over time. Additional note: The PTFE-lined scintillation vial has to
be used for organic solvents. Never leave the vial with toluene in the incubat-
64 3 Chip Platforms and Chip Surface Treatments
For enzyme printing, amine-reactive surfaces may deactivate the enzymes printed.
Thus, the surface of acid-cleaned glass slides is typically changed to hydrophobic.
A commonly used method is to coat the surface of the glass slides with hydrophobic
silicone precursors such as MTMOS.
1. Prepare a fresh MTMOS-HCl sol solution by mixing 8 mL of MTMOS with
3.2 mL of 5 mM HCl, followed by vortex for 2 min and sonication for 10 min.
Note: The hydrolysis of MTMOS by HCl is an exothermic reaction. So be
careful not to burn skin. The mixture becomes transparent single phase after
vortex mixing (initially two phases). Always prepare fresh MTMOS-HCl sol
solution and use it immediately. Leaving the MTMOS-HCl sol solution longer
than 30 min before spin coating cause uneven coating.
2. Immediately before spin coating, mix 11.2 mL of the MTMOS-HCl sol solution
with 8 mL of potassium phosphate buffer solution (25 mM, pH 8) and use the
mixture within 15 min. Note: Upon addition of the phosphate buffer solution,
gelation of hydrolyzed MTMOS begins. Therefore, use the mixture immedi-
ately. Do not use the mixture when the solution becomes turbid, causing uneven
coating.
3. While spinning the acid-cleaned glass slide at 3000 rpm for 30 s, place 1.5 mL
of the mixture onto the slide.
4. Remove any excess MTMOS on the back of the slide using acetone-soaked
paper towel. Note: Make sure that rainbow-like refraction patterns are seen on
the surface after spin coating. If not, then the MTMOS coating is unsuccessful,
3.3 Protocols 65
presumably because either the acid-cleaned glass slide is too dirty or the
MTMOS-HCl-phosphate buffer mixture is too old.
5. To complete MTMOS gelation, dry the slides overnight at room temperature in
a sterile petri dish and store until use.
1. Allow DMSO (component B) and one vial of reactive FITC dye (component A)
from Life Technologies (FluoReporter® protein labeling kit) to warm to room
temperature immediately before starting the reaction.
66 3 Chip Platforms and Chip Surface Treatments
2. Add 50 μL of DMSO to the vial of reactive FITC dye. Note: Protect the reac-
tive FITC dye from light.
3. Prepare 200 mL of 50 mM potassium phosphate buffer (pH 8 from Life
Technologies) by diluting with de-ionized distilled water. Note: The buffer solu-
tion should not contain any ammonium ions or primary amines because the dye
is reactive with amine groups. Thus, Tris or glycine buffers are unsuitable.
4. Add 50 μL of reactive FITC dye in 200 mL of phosphate buffer and mix well
with a magnetic stir bar. Any remaining reactive dye stock solution should be
discarded.
5. Place amine-functionalized slides and acid-cleaned glass slides (as a control) in
a removable glass rack, and then immerse the rack in a Wheaton glass dish
filled with 100 % acetonitrile (ACN) for 5 min.
6. Rinse the slides in series for 5 min with an equal volume mixture of ACN and
water, and de-ionized distilled water to wet the surface.
7. Immerse the slides in the reactive FITC dye solution and incubate them for 1 h
with magnetic stirring at room temperature. Note: Do not sonicate the slides in
the reactive FITC dye solution because sonication can cause non-specific
binding of the dye on the surface. While incubating the slides, protect the
fluorescent dye solution from light by wrapping it up in aluminum foil.
8. Rinse the slides three times in de-ionized distilled water for 1 h each with magnetic
stirring to remove unbound dye.
9. Rinse the slides in acetone to remove excess water and wrap them up in alu-
minum foil. Note: Be careful not to scratch the surface of the slides with
aluminum foil.
10. Measure fluorescent intensity on the slides at 494 nm (excitation) and 518 nm
(emission).
1. Rinse the amine-functionalized slides in series for 5 min with 100 % acetonitrile
(ACN), an equal volume mixture of ACN and water, and de-ionized distilled
water to wet the surface. Sterilize de-ionized distilled water with Nalgen MF75
series sterile disposable tissue culture filter unit (0.2 μm PES filter).
2. Prepare 200 mL of 5 % (v/v) of glutaraldehyde in sterile Dulbecco’s phosphate-
buffered saline (PBS) without CaCl2 & MgCl2.
3. Immerse the amine-functionalized slides in the glutaraldehyde solution and soni-
cate them for 1 h at room temperature. Discard the used glutaraldehyde solution
after incubation.
4. Rinse the slides twice by dipping in de-ionized distilled water for 5 min to
remove unbound glutaraldehyde. Use sterile de-ionized distilled water.
5. Rinse the slides in acetone and dry them by blowing nitrogen. Note: Use the
slides as soon as prepared because of instability of glutaraldehyde.
3.3 Protocols 67
ECM proteins can be covalently attached on the reactive surfaces of glass slides
such as PS-MA-coated slides and glutaraldehyde-functionalized slides (Figs. 3.12
and 3.13).
HC
C
O O
H O
HC
Si O Si (CH2)3 N C CH C
Si OH O O
HO C CH
CH CH2
Si OH O
CH CH2
Si OH
NH2
OCH3
H3CO Si (CH2)3 NH2 Hydrophobic interaction
OCH3
O
HC
O O O C OH
H
Si O Si (CH2)3 NH2 Si O Si (CH2)3 N C HC
CH C NH Collagen
O O
O
HO C CH
CH CH2
O
CH2 CH CH CH
CH CH2
O C CO y
x O
Si
O
Si
O
Si
O O
(MeOH+HCl) & H2SO4 5% HC (CH2)3 CH
Si OH O O
H
Si O Si (CH2)3 N C (CH2)3 CH
Si OH
O
Si OH
OCH2CH3
10% H3CH2CO Si (CH2)3 NH2 NH2
OCH2CH3
O O H H
Si O Si (CH2)3 NH2 Si O Si (CH2)3 N C (CH2)3 C N
O O
1. Prepare ECM protein solutions (40 μg/200 μL) in sterile PBS (1×) and place
them on ice. Typical ECM proteins used are as follows: Matrigel™, collagen I,
fibronectin, laminin, vitronectin, etc.
2. Print 60 nL of ECM protein solutions on the reactive surfaces of glass slides
using S+ MicroArrayer.
3. Incubate the slides for 1 h at room temperature in a humid incubation chamber
to prevent evaporation of water (Fig. 3.14).
4. After drying the slides at room temperature in a biosafety cabinet, keep them in
a sterile petri dish for long-term storage. It may be stored at 2–8 °C for up to
several weeks under sterile conditions.
5. Before adding cell suspensions (e.g., Hep3B, MCF7, and A293 cells), rinse the
slides in sterile PBS twice with shaking for 30 min at 100 rpm to remove unbound
ECM proteins.
6. Culture the cells in proper growth media for 2–3 days in a CO2 incubator at
37 °C to prepare cell patterns or confluent cell monolayers.
3.4 Summary
Glass and polystyrene are both suitable substrates for cellular microarrays, but both
substrates require chemical modifications to ensure proper adhesion of cells or
hydrogels. In the case of plastic chips made of polystyrene, it involves coating with
amphiphilic, water-insoluble polymers. For glass slides, this means chemical
modification with silanes, or spin-coating with amphiphilic polymers. In both cases,
these only serve as preliminary steps to ensure adequate biocompatibility and
cell/hydrogel attachment, as subsequent steps are usually necessary to generate
microarrays. Nevertheless, these steps are necessary to create surfaces that can be
used for biological assay development with living cells.
References 69
References
1. Lee, M.-Y., Kumar, R. A., Sukumaran, S. M., Hogg, M. G., Clark, D. S., & Dordick, J. S. (2008).
Three-dimensional cellular microarray for high-throughput toxicology assays. Proceedings of
the National Academy of Sciences, 105(1), 59–63. doi:10.1073/pnas.0708756105.
2. Ramsey, W., Hertl, W., Nowlan, E., & Binkowski, N. (1984). Surface treatments and cell
attachment. In Vitro, 20(10), 802–808.
3. Curtis, A. S. G., Forrester, J. V., McInnes, C., & Lawrie, F. (1983). Adhesion of cells to poly-
styrene surfaces. The Journal of Cell Biology, 97(5I), 1500–1506. doi:10.1083/jcb.97.5.1500.
4. Amstein, C. F., & Hartman, P. A. (1975). Adaptation of plastic surfaces for tissue-culture by
glow-discharge. Journal of Clinical Microbiology, 2(1), 46–54.
5. Hudis, M. (1974). Plasma treatment of solid materials. In R. H. Hansen & A. T. Bell (Eds.),
Techniques and applications of plasma chemistry (p. 147). New York: John Wiley and Sons.
6. Seo, J. H., Shin, D. S., Mukundan, P., & Revzin, A. (2012). Attachment of hydrogel micro-
structures and proteins to glass via thiol-terminated silanes. Colloids and Surfaces B:
Biointerfaces, 98, 1–6. doi:10.1016/j.colsurfb.2012.03.025.
7. Carvalho, R. R., Pujari, S. P., Lange, S. C., Sen, R., Vrouwe, E. X., & Zuilhof, H. (2016). Local
light-induced modification of the inside of microfluidic glass chips. Langmuir, 32(10), 2389–
2398. doi:10.1021/acs.langmuir.5b04621.
8. Luo, N., Zhong, H., Yang, M., Yuan, X., & Fan, Y. (2016). Modifying glass fiber surface with
grafting acrylamide by UV-grafting copolymerization for preparation of glass fiber reinforced
PVDF composite membrane. Journal of Environmental Sciences, 39, 208–217. doi:10.1016/j.
jes.2015.11.010.
9. Cheawchan S, Uchida S, Sogawa H, Koyama Y, & Takata T. Thermotriggered catalyst-free
modification of a glass surface with an orthogonal agent possessing nitrile N-Oxide and
masked Ketene functions. Langmuir 2015;32(1): 309–315. doi:10.1021/acs.langmuir.5b03881.
(Fig. 1).
10. Tang, J., Peng, R., & Ding, J. (2010). The regulation of stem cell differentiation by cell-cell
contact on micropatterned material surfaces. Biomaterials, 31(9), 2470–2476. doi:10.1016/j.
biomaterials.2009.12.006.
11. User’s manual for PDC-001-HP (115V) or PDC-002-HP (230V) high power expanded plasma
cleaner (and optional PDC-FMG (115V) or PDC-FMG-2 (230V) PlasmaFlo). Ithaca, NY;
2015.
12. Cras, J. J., Rowe-Taitt, C. A., Nivens, D. A., & Ligler, F. S. (1999). Comparison of chemical
cleaning methods of glass in preparation for silanization. Biosensors & Bioelectronics, 14(8–9),
683–688. doi:10.1016/S0956-5663(99)00043-3.
13. No Title. Retrieved from http://www-schreiber.chem.harvard.edu/home/protocols/SMP.htm.
Chapter 4
Biological Sample Printing
Contents
4.1 Introduction...................................................................................................................... 71
4.2 Materials.......................................................................................................................... 73
4.3 Preparation of Stock Solutions......................................................................................... 74
4.4 Daily Operation of the S+ MicroArrayer for Dispensing Biological Samples................ 76
4.5 Sample Printing Protocols............................................................................................... 80
4.5.1 Enzyme Printing for Metabolism-Induced Drug Toxicity Assays....................... 80
4.5.2 Compound Printing.............................................................................................. 82
4.5.3 Cell Printing......................................................................................................... 85
4.5.3.1 Preparation of Cell Suspension in Growth Medium........................... 85
4.5.3.2 Cell Printing for 2D Monolayer Culture on the Micropillar Chip...... 86
4.5.3.3 Cell Printing in Alginate for 3D Cultures........................................... 87
4.5.3.4 Cell Printing in Matrigel for 3D Cultures........................................... 90
4.5.3.5 Cell Printing in a Mixture of Alginate
and Matrigel for 3D Cultures.............................................................. 93
4.5.3.6 Cell Printing in a Mixture of Alginate and Fibrinogen
for 3D Cultures................................................................................... 95
4.5.3.7 Cell Printing in PuraMatrix for 3D Cultures...................................... 96
4.5.4 Virus Printing....................................................................................................... 98
4.5.4.1 Measurement of Viral Titer in a 96-Well Plate................................... 98
4.5.4.2 Adenoviral Transduction on the Micropillar/Microwell Chip............ 99
4.6 Inspection of Cells Printed on the Micropillar Chips
Using a Bright-Field Microscope..................................................................................... 101
4.7 Coefficient of Variation (CV) and Z’ Factor for Assay Validation.................................. 102
4.8 Summary.......................................................................................................................... 103
References................................................................................................................................. 103
4.1 Introduction
Microarray spotters equipped with solenoid valves are capable of printing a wide
range of biological samples, including reagents, growth media, compounds, hydro-
gels, genes, proteins, viruses, and cells, for biochemical and cell-based assays.
Solenoid valve-driven bioprinting has clear advantages over other printing technologies
in terms of controlling sample volume dispensed and flexibility in the type of sam-
ples printed. Unlike robotic liquid dispensers that are more robust due to relatively
large orifice sizes and dispensing volumes, microarray spotters must take precau-
tions to prevent the clogging of solenoid valves and ceramic tips due to the extremely
small sample volume printed. Therefore, it is essential to avoid dust and large pre-
cipitates (e.g., compound precipitates and filamentous microbes), test the viscosity
of the material printed, and understand the mechanism of gelation for hydrogels
used for cell encapsulation. In general, temperature-sensitive hydrogels such as
Matrigel have a high risk of forming a gel spontaneously inside tubes, ceramic tips,
and solenoid valves with a slight change in temperature. Therefore, a lower tem-
perature must be maintained constantly across the chip-loading deck and the dis-
pensing head. The clogging may result in replacement of tubes, solenoid valves, and
ceramic tips, which can be an expensive affair. Gelation of hydrogels used for 3D
cell culture has to be occurred in two steps to avoid clogging. Typically, cross-
linkers/initiators are printed first on the micropillar/microwell chip platform and
then hydrogels containing cells are dispensed on the top to form a gel. Therefore,
ionic crosslinking (e.g., alginate with CaCl2, PuraMatrix with salts), affinity/cova-
lent bonding (e.g., functionalized polymers with streptavidin and biotin), photopo-
lymerization (e.g., methacrylated alginate with photoinitiators), and biocatalysis
(e.g., fibrinogen with thrombin) are favorable mechanisms of gelation. Developing
proper surface chemistry to attach cell spots in hydrogels robustly on the surface of
the micropillar/microwell chip is also critical. Covalent bonding (e.g., poly(maleic
anhydride-alt-1-octadecene) and poly-l-lysine), affinity (e.g., streptavidin and bio-
tin), and ionic interaction (poly-l-lysine and negatively charged alginate) are com-
monly introduced to enhance spot attachment on the chip. Viscosity of the material
printed is another important factor that affects the performance of printing, and
valve/tip clogging and rinsing. In general, biological samples with low viscosity
such as 10–30 centipoise (equivalent to 1 % or lower alginate in distilled water) are
preferred. Samples with high viscosity cannot be printed with solenoid valves and
are difficult to remove from tubes, solenoid valves, and tips by rinsing with water.
Therefore, it has to be diluted properly with either solvents or cell culture media to
ensure reproducible printing. In case of cell printing and encapsulation for 3D cul-
ture, maintaining cell suspension in hydrogels while printing is important to avoid
solenoid valves and tips clogging. In general, over crowded cells (typically seeding
density higher than 10 million cells/mL) can result in clogging issues. In addition,
cells tend to settle down quicker in a low viscosity solution. Finally, mechanical
strength of hydrogels over time needs to be considered to support long-term cell
culture. Peptide-based hydrogels (e.g., fibrinogen, Matrigel, and PuraMatrix) tend
to lose their strength over time due to degradation by matrix metalloproteinases
(MMPs) secreted by many mammalian cells. These MMPs are responsible for
hydrogel degradation and eventual spot detachment. To sustain cell spots for a
longer time and minimize spot detachment due to degradation, peptide-based
hydrogels can be mixed with more stable hydrogels such as alginate. When mixed,
the resulting hydrogel has to be transparent with minimal background fluorescence
4.2 Materials 73
and should not interfere high-content imaging (HCI) assays. The protocols provided
in this chapter will give researchers a guidance towards multiplex, 3D-cell based
assays on the micropillar/microwell chip platform.
4.2 Materials
1. Open the air cylinder (or in-house air valve) and maintain the pressure of com-
pressed air at 100 psi.
2. Turn on the chiller, the external switch in the utility body, the computer, and the
monitor.
3. Prior to running the ‘ezAOI’, push ‘Reset button’ and reset XYZ coordinates.
Note: This step is essential to avoid malfunctioning of the microarray spotter.
Do not skip this step!
4. Run the ‘ezAOI’ software (Fig. 4.1). Note: Make sure that no obstacles are on
the work place to avoid the robotic arm crashing.
5. Go to ‘Water Alcohol Change’ in the ‘User Operation’ box (Fig. 4.2).
6. Click ‘Release’ in the ‘Air pressure’ box to release air pressure in the pressure
bottles.
Fig. 4.1 The main screen of ezAOI. The Command window has four groups—Equipment setup,
Program, User operation, and Work
4.4 Daily Operation of the S+ MicroArrayer for Dispensing Biological Samples 77
7. Turn the orange knob connected to the pressure bottles from ‘S’ to ‘O’ position,
which releases the pressure inside.
8. Open the lid of the pressure bottles and fill the pressure bottles with distilled
water and 70 % ethanol.
9. Close the lid of the pressure bottles carefully. Note: Make sure that the ‘O’
ring is properly placed before placing the lid.
10. Turn the orange knob of the pressure bottles from ‘O’ to ‘S’ position, which
close the pressure bottles completely.
11. Click ‘Initialize’ of the ‘Air pressure’ box in the ‘Water Alcohol Change’ box
to apply air pressure in the bottles.
12. Click the ‘Start’ button in the ‘Daily Washing’ box. Note: Washing tubes and
solenoid valves with alcohol is essential to remove air bubbles prior to sample
dispensing. After alcohol washing, thorough rinsing of tubes and solenoid
valves with water is necessary to avoid enzyme deactivation or cell death due
to remaining alcohol. An extensive rinse of tubes and solenoid valves with
ice-cold water is critical before and after printing Matrigel solutions.
13. Check a streamline of water while washing. Note: If the streamline is deflected
from the vertical or water is beaded up at the tip end, clean the ceramic tip
with sonication.
14. Select the ‘Work’ window (Fig. 4.3).
78 4 Biological Sample Printing
15. Click the ‘File Open’ button, and select the desired work file.
16. Enter the offset values in the X and Y directions to designate the location of
samples in 96-wells. The exact location of sample wells will be appeared in the
schematic of the well plate in the ‘Display’ window (Fig. 4.4).
17. Click the ‘Chip Loading’ button, and place the chips on the chilled chip deck
shortly before sample printing and select the location of chips in the ‘Display’
window onto which samples are dispensed (Fig. 4.4).
18. Add proper amounts of samples based on aspiration and sacrificial volumes
shown in the ‘Display’ window, in the designated wells, and place the well plate
on the deck. Note: Always add 20 μL more to the aspiration and sacrificial
volumes in the 96-well plate in case of evaporation.
19. Click the ‘Run’ button.
20. To avoid potential contamination issues after sample printing, rinse tubes and
solenoid valves with ethanol and water thoroughly by running ‘Daily Washing’.
21. Close ‘ezAOI’ software.
80 4 Biological Sample Printing
22. Turn off the computer, the monitor, the external power switch in the utility
body, and the chillers. Note: Make sure to turn off the chillers to avoid excess
water condensation on the chip-loading deck and the dispensing head, caus-
ing a short circuit by water.
23. Close the air cylinder (or in-house air valve). Note: Do not close the air c ylinder
while ‘ezAOI’ is on.
Fig. 4.5 (a) A layout of the MetaChip (432 spots/chip) for in-situ drug metabolism. Specifically,
region A contains no enzyme as a test compound only control, regain B contains a mixture of
human Cytochrome P450 isoforms, region C contains a mixture of human phase II drug-
metabolizing enzymes and P450 enzymes (i.e., all enzyme mixture), and region D contains human
liver microsome. (b) A layout of compounds printed in the MetaChip (432 spots/chip). Each dif-
ferent compound is printed in sections 1–6 of the MetaChip, and the concentrations (C1–C6) vary
within each mini-array. The grey dots represent the microwells
At least 2 cycles of running the wash program with cold water is necessary to
cool it down.
5. Prepare cold metabolizing enzyme solutions in Matrigel in a 96-well plate on ice
(see Table 4.1 for details). Note: Always use the same kind of the well plates to
avoid tip crashing due to Z height difference in wells. Always prepare fresh
enzyme-Matrigel mixtures on ice and use them immediately.
82 4 Biological Sample Printing
Test compounds can be printed in the MetaChip with drug metabolizing enzymes
for assessing metabolism-induced toxicity or in the microwell chip (with no
enzymes) for measuring parent compound toxicity. Typical printing volume of com-
pounds in growth media on the chip is 800–950 nL. It is critical to minimize the use
of organic solvents in the compound solutions. Typically, less than 0.5 % of DMSO
4.5 Sample Printing Protocols 83
A1 A2 A3 A4 A5 A6
DMSO Compound
alone in DMSO
A1 A2 A3 A4 A5 A6
B1 B2 B3 B4 B5 B6
C1 C2 C3 C4 C5 C6
D1 D2 D3 D4 D5 D6
E1 E2 E3 E4 E5 E6
F1 F2 F3 F4 F5 F6
(A)
A1 A2 A3 A4 A5 A6
B1 B2 B3 B4 B5 B6
C1 C2 C3 C4 C5 C6
D1 D2 D3 D4 D5 D6
E1 E2 E3 E4 E5 E6
F1 F2 F3 F4 F5 F6
(B)
Fig. 4.6 (a) Preparation of a compound-in-DMSO plate in a 96-well plate by fourfold serial dilu-
tion. The 96-wells A6, B6, C6, D6, E6, and F6 contain 40 μL of compound stocks in DMSO and
the remaining 96-wells containing 30 μL of DMSO. For fourfold serial dilution, 10 μL of com-
pounds in DMSO are transferred to adjacent 96-wells and mixed well with a multi-channel pipette
sequentially until the last 96-wells with DMSO alone remains. Thus, the 96-wells A1, B1, C1, D1,
E1, and F1 containing DMSO alone are used as controls. (b) Preparation of a compound-in-growth
medium plate in a 96-well plate. For 200-fold dilution of compounds in growth media, dispense
298.5 μL of growth media in the 96-wells A1–F6, transfer 1.5 μL of serially diluted compounds in
the compound-in-DMSO plate to the 96-well plate containing growth media with a multi-channel
pipette, and mix the solutions well by aspirating and dispensing at least 20 times. Samples in the
96-wells A1, B1, C1, D1, E1, and F1 are transferred first, and then sequentially from low to high
concentrations to avoid compound carry-over
4.5 Sample Printing Protocols 85
It is very important to prepare well suspended cells in growth media prior to mixing
with hydrogels for cell printing. Clumpy cells can clog solenoid valves and ceramic
tips, resulting in unreliable printing problems. Always make sure to resuspend cells
in the 96-well plate manually with a pipette before sample aspiration for uniform
printing.
86 4 Biological Sample Printing
1. Carefully remove the medium from a tissue culture flask with approximately
80 % confluent monolayer of mammalian cells.
2. Rinse the flask once with DPBS. Add DPBS slowly from the side to avoid
detaching the cells. Use 12 mL of DPBS for a T75 flask.
3. Aspirate out DPBS.
4. Add 1–2 mL of 0.05 % trypsin in the T75 flask, uniformly coat the cell mono-
layer with the trypsin solution, and incubate for 1–2 min at 37 °C.
5. Inspect the flask and ensure complete detachment of the cells by gently tapping
the side of the flask with the palm of your hand.
6. Apply 5 mL of growth medium pre-warmed at 37 °C to the flask.
7. Aspirate and dispense the growth medium containing cells rigorously at least
20 times to break apart big cell clumps. Note: The trypsin step and this step are
critical to prepare good cell suspension as clumpy cells can clog solenoid
valves and ceramic tips.
8. Gently rotate the flask to suspend the cells in growth medium.
9. Transfer the detached and suspended cells to a 15 mL conical tube.
10. Centrifuge the tube at 1200 rpm (typically 200 g) for 4 min to pellet the cells.
11. Aspirate out the entire supernatant gently without disturbing the cell pellet.
12. Add 2 mL of a complete growth medium supplemented with 10 % FBS and 1 %
Pen/Strep and resuspend the cells thoroughly with a 1 mL pipette.
13. While cells are in good suspension, take 5 μL of the cell suspension out and mix
it with 495 μL of growth media in an Eppendorf tube (1.5 mL) for cell counting
at 100-fold dilution.
14. After well mixing with a pipette, take 75 μL of cell suspension out from the
Eppendorf tube and load it in a Moxi Z cassette.
15. Insert the cassette in a Moxi Z mini automated cell counter to measure cell
density in the number of cells per milliliter. Actual cell density is 100-fold
higher than this density because of the dilution factor. Note: Do not vortex the
cell suspension for mixing as mammalian cells are fragile.
16. Dilute the cell suspension with growth media at a desired seeding density (typi-
cally, 6–10 million cells/mL) and then mix it with a hydrogel solution. The final
cell seeding density is typically in the range of 1–4 million cells/mL, varied
depending on printing volume on the chip and doubling time of the cells printed.
The required volume of cell suspension can be calculated from the program
(ezAOI) in S+ MicroArrayer by selecting an appropriate work file and the num-
ber of chips necessary.
1. Coat the micropillar chip with 2 mL of 0.01 % (w/v) PMA-OD in ethanol, and
let it dry at room temperature at least for 2–3 h (see the Sect. 3.3.2 in Chap. 3).
2. Prepare 0.0033 % poly-l-lysine (PLL) by mixing 0.01 % PLL stock (Sigma-
Aldrich) with sterile deionized water at 1:2 ratio to obtain a final concentration
of 0.0033 % PLL. Note: In order for robust cell attachment onto the micropil-
lar chip, in addition to PLL, we may need collagen or laminin coating for
some cells.
3. Print 60 nL of 0.0033 % poly-l-lysine (PLL) on the micropillar chip using a
work file and dry it for 2–3 h. Note: Coating the micropillar chip with biocom-
patible polymer solutions is critical for robust attachment of cells on the
surface.
4. Prepare cell suspension in the complete growth medium at a desired seeding
density (1–6 million cells/mL).
5. Print 950 nL of the cell suspension in the microwell chips. Note: It is important
to resuspend the cells in the growth medium immediately before printing.
Printing cell suspension at too high seeding density may cause clogging of
solenoid valves and ceramic tips.
6. Sandwich the micropillar chip with polymer coating with the microwell chip
containing cell suspension.
7. Place the sandwiched chips with the microwell chip on the top in a humidified
petri dish with 10 mL of distilled water so that the cells can be settled on the
surface of the micropillars.
8. To robustly attach the cells on the surface of the micropillars, incubate the sand-
wiched chips in the humidified petri dish in a 5 % CO2 incubator at 37 °C for
6 h. Note: It is extremely important to prevent water evaporation during cell
incubation as drying can result in cell death.
9. After 6 h incubation, flip the sandwiched chips around to place the micropillar
chip on the top, which may prevent leaching the growth medium out.
10. Incubate the sandwiched chips overnight for cell growth.
11. After robust cell attachment on the micropillar chip, change the growth medium
every 1–2 days by printing fresh growth medium in the microwell chip, discard-
ing the microwell chip containing old growth medium, and sandwiching the
freshly prepared microwell chip. Note: The cells are cultured on the micropillar
chip until 80–90 % confluency.
Since 2D-cultured cells rapidly lose phenotypic properties and functions of tissues
and may not accurately emulate the microenvironment and cellular architecture
found in vivo, 3D cell culture strategies have been developed in efforts to foster
retention of phenotypic tissue functions. We have successfully demonstrated human
cell cultures in 3D on the micropillar/microwell chip platform using various hydro-
gels, including alginate, PuraMatrix™, fibrinogen, Matrigel, Geltrex, and photo-
crosslinkable, methacrylated alginate. In particular, alginate is an ideal printable
scaffold for tissue engineering because it is a naturally-derived hydrogel that is non-
toxic, non-immunogenic, and non-adhesive in nature [5]. Alginate is also able to
88 4 Biological Sample Printing
deliver soluble signaling molecules to cells, act as a support structure for cell growth
and function, and provide space filling for future tissue ingrowth after hydrogel
degradation [5]. Cells can easily be encapsulated and remain viable in alginate
hydrogels, and it has been widely used in a variety of tissue engineering applica-
tions [5]. Alginate is a negatively charged biopolymer which forms a gel when
interacting with a divalent cation such as Ca2+, Ba2+, Mg2+, etc. Some divalent cat-
ions such as Ba2+ can be relatively toxic to cells, thus requiring additional rinsing of
cells to remove excess Ba2+ ions. Alginate can be functionalized with methacrylate
groups for photopolymerization [5, 6]. Photocrosslinkable alginate has been used as
a biodegradable hydrogel system with independently adjustable physical, cell adhe-
siveness, and bioactive factor binding properties [7, 8]. Optimization of photoinitia-
tors and their concentrations are necessary to minimize basal toxicity of cells
printed. The protocol of cell printing in alginate using S+ MicroArrayer is summa-
rized below.
1. Coat the micropillar chip with 2 mL of 0.01 % (w/v) PMA-OD in ethanol and
dry it at room temperature at least for 2–3 h.
2. Prepare a mixture of 0.0033 % PLL and 25 mM CaCl2 by mixing 0.01 % PLL
with 37.5 mM CaCl2 in distilled water at 1:2 ratio. Note: For human cell lines,
a mixture of 0.0033 % PLL and 16 mM BaCl2 can be used to improve spot
attachment. BaCl2 is much stronger crosslinker, but more toxic against some
primary human cells and delicate cells.
3. Print 60 nL of the PLL-CaCl2 mixture on the micropillar chip using a work file
(60 nL PLL_CaCl2) and dry it for at least 2–3 h.
4. Print 950 nL of growth medium in the microwell chip placed on the chilling chip
deck at 4–12 °C using a work file (medium_6tip_6block_950nL). Note: Turn on
the chiller at least 1 h before sample printing.
5. Immediately after printing, incubate the microwell chips with growth medium in
a humid chamber with water to avoid water evaporation (Fig. 4.7).
6. Prepare 1–4 million cells/mL in 0.75 % alginate by mixing 1 mL of 2–8 million
cells/mL, 0.5 mL of 3 % alginate in distilled water, and 0.5 mL of growth medium
using a 1 mL pipette. Note: Do not vortex the mixture as mammalian cells are
Fig. 4.8 Culture of cells on sandwiched chips in (a) a humid incubation chamber with 25 mL
water and (b) a petri dish with 10 mL water
90 4 Biological Sample Printing
14. Inspect the sandwiched chips under a bright field microscope with 4× or 10×
objective lens for spot attachment and uniform cell printing. Note: Make sure
to avoid water evaporation during chip inspection. Do not leave out the
sandwiched chips longer than 2 min for chip inspection. Water evaporation
will change the concentration of salts in growth medium, thus influencing
osmotic pressure inside cells and reducing cell viability and growth.
15. Incubate the sandwiched chips in the humidified petri dish placed in a 5 % CO2
incubator at 37 °C for 3D cell culture.
16. Change growth medium every 1–2 days by printing fresh growth medium in the
microwell chip, discarding the microwell chip containing old growth medium,
and sandwiching the freshly prepared microwell chip.
Matrigel consists of laminin, collagen IV, entactin, heparin, and entactin which is pro-
duced by Engelbreth-Holm-Swarm (EHS) mouse sarcoma cells. It has been widely
used in cell growth, differentiation, angiogenesis, and tissue vascularization, since it
resembles cellular microenvironment in various tissues [9, 10]. Although Matrigel can
be used for cell printing and encapsulation on the chip platform, extra caution has to be
exercised as Matrigel is a temperature-sensitive hydrogel that forms a gel at a tempera-
ture ranging from 24 to 37 °C [11]. Therefore, it is necessary to maintain a low tem-
perature on the chip-loading deck and in the dispensing head while printing Matrigel to
avoid premature gelation and clogging in solenoid valves and ceramic tips. Typically,
cells are mixed with cold Matrigel and printed immediately onto the micropillar chip.
The printed cells in Matrigel can be gelled at 37 °C within 30 min [12]. Specific
example of primary hepatocyte printing in Matrigel is presented below.
ing the cap and then retightening it, and remove the paper label so that the ice
crystal melting can be easily observed during the thawing process.
3. Immerse the cryo-vial containing hepatocytes in a 37 °C water bath and gently
shake the cryo-vial in water until the last ice crystals floated to the top of the
cryo-vial. Note: Do not take the vial out of water during the thawing process
to check ice crystal melting. Always check it in water. The whole thawing
process typically takes about 3 min.
4. Immediately pour the melted hepatocytes from the cryo-vial into 50 mL of the
CHRM™ medium.
5. Rinse the cryo-vial three times with 1 mL of the CHRM™ medium to recover
hepatocytes attached on the wall.
6. Centrifuge the hepatocyte suspension in the CHRM™ medium at 100 g (RCF)
and 20 °C for 10 min with the slowest brake and acceleration settings available on
the centrifuge. Note: This is a critical step as primary hepatocytes are delicate.
7. Aspirate out the supernatant into a waste bottle, leaving the hepatocyte pellet at
the bottom.
8. Pour 3 mL of the thawing/plating medium onto the hepatocyte pellet and resus-
pend the pellet by gently swirling the tube or by tapping on the hood surface or
by dragging the bottom of the tube over corrugated surface. Note: Do not use a
vortex or a pipette to make well suspension of hepatocytes as they are very
fragile. Do not shake the hepatocyte suspension.
9. Split the thawing/plating medium containing suspended hepatocytes into two in
Eppendorf tubes using a 1 mL pipette.
10. Centrifuge the hepatocyte suspension in thawing/plating medium again under
the conditions stated above.
11. Aspirate out the supernatant into a waste bottle, leaving the hepatocyte pellet at
the bottom.
12. Add 1.5 mL of the thawing/plating medium onto the hepatocyte pellet and
resuspend the resulting hepatocyte pellet by gentle swirling. Note: Primary
hepatocytes are extremely fragile. Do not vortex, shake, or aspirate/dispense
hepatocytes to make the suspension.
13. Combine two hepatocyte suspensions into one and centrifuge the combined
hepatocyte suspension in thawing/plating medium under the conditions stated
above.
14. Aspirate out the supernatant into a waste bottle, leaving the hepatocyte pellet at
the bottom.
15. Add 300 μL of the thawing/plating medium onto the hepatocyte pellet, resus-
pend the resulting hepatocyte pellet by gentle swirling, and then store the
hepatocyte suspension (8 million cells/mL) on ice until use.
16. To count the total number of hepatocytes, remove 10 μL of the hepatocyte
suspension, add 490 μL of the thawing/plating medium, and measure the cell
number in suspension using an automated cell counter.
17. To measure the viability of hepatocytes, remove 10 μL of the hepatocyte sus-
pension, add 40 μL of the thawing/plating medium, mix the diluted hepatocyte
suspension with 50 μL of trypan blue, and place 10 μL of the hepatocyte-trypan
blue mixture on a hemocytometer.
92 4 Biological Sample Printing
11. Sandwich the micropillar chip (typically 532 spots/chip, 1.5 mm pillar-to-pillar
distance) with the shallow-well plate containing the cold Matrigel-DPBS
mixture and then dry Matrigel-DPBS spots on the micropillar chip for 1 h in a
sterile petri dish with a lid slightly open to yield flat Matrigel layers.
12. Prepare a suspension of hepatocytes in Matrigel on ice by mixing 300 μL of the
hepatocyte suspension with 300 μL of Matrigel so that the final concentration
of cells will be 4 × 106 cells/mL. Note: To prepare well-suspended hepatocytes
in Matrigel, gently mix them on ice by aspirating and dispensing for 10 times.
Don’t mix cell suspension with Matrigel by vortex as some cell membranes
are very fragile.
13. Print the cold Matrigel solution containing hepatocytes (532 spots/chip, 50 nL
per spot) atop each Matrigel-DPBS spot at 4 °C. Note: Put the 96-well plate
with the cold Matrigel-hepatocyte mixture back on ice immediately after aspi-
ration, or chill the 96-well plate deck by turning on the chiller and setting the
temperature at 4 °C.
14. Incubate the micropillar chip with hepatocytes in the black humid chamber
with 700 μL water in each well at 37 °C for 20 min, resulting in gelation of the
Matrigel matrix (Fig. 4.7). Note: Ensure the Matrigel gelling condition in
your lab by adjusting the incubation time at 37 °C.
15. If necessary, add freshly Matrigel solution containing suspended hepatocytes in
the round-bottom 96-well plate to make sure well suspension of hepatocytes
while spotting. After printing every chips (i.e., after every 5–10 min of spot-
ting), add freshly Matrigel solution containing suspended hepatocytes in the
well repeatedly. Note: Remind that hepatocytes are gradually settling down
while spotting. For uniform cell spotting, keep re-suspending hepatocytes.
After cell spotting, operate the daily washing program at least twice with ice-
cold water to make sure that no Matrigel remains in tubes, solenoid valves,
and ceramic tips.
16. After incubation of the micropillar chip for Matrigel gelation, stamp the micropillar
chip with hepatocytes onto the microwell chip with 950 nL of the thawing/plating
medium, place the stamped chips in the gas-permeable chamber, and then incubate
it in a 5 % CO2 incubator at 37 °C for 2 h prior to compound toxicity tests.
17. Prepare compound solutions in the maintenance medium, print the medium in
the microwell chip, stamp the micropillar chip with hepatocytes onto the
microwell chip with compound solutions, and then incubate the stamped chips
in a 5 % CO2 incubator for 24 h for acute toxicity assays.
18. Stain hepatocytes on the micropillar chip with a Live⁄Dead® viability/cytotoxicity
kit from Life Technologies.
4.5.3.5 C
ell Printing in a Mixture of Alginate and Matrigel for 3D
Cultures
Matrigel can be mixed with alginate to enhance cell viability and growth in alginate,
while avoiding temperature-induced gelation. Thus, the concentration of Matrigel
in alginate is substantially lower than its counterpart, Matrigel alone. The general
protocol of cell printing in a mixture of alginate and Matrigel is presented below.
94 4 Biological Sample Printing
1. Coat the micropillar chip with 2 mL of 0.01 % (w/v) PMA-OD in ethanol and
dry it at room temperature at least for 2–3 h.
2. Prepare a mixture of 0.0033 % PLL and 25 mM CaCl2 by mixing 0.01 % PLL
with 37.5 mM CaCl2 in distilled water at 1:2 ratio.
3. Print 60 nL of the PLL-CaCl2 mixture on the micropillar chip using a work file
(60 nL PLL_CaCl2) and dry it for at least 2–3 h.
4. Print 950 nL of growth medium in the microwell chip placed on the chilling
chip deck at 4–12 °C using a work file (medium_6 tip_6 block). Note: It is
particularly important to maintain low temperature when printing Matrigel.
Turn on both chillers connected to the chilling chip deck, the dispensing
head, the 96-well plate deck, and the water/alcohol containers at least 1 h
before sample printing. Wait until the temperatures reach 4–12 °C. Matrigel
is gelled at 37 °C.
5. Immediately after printing, incubate the microwell chips with growth medium
in a humid chamber with water to avoid water evaporation (Fig. 4.7).
6. Take out one aliquot of 500 μL Matrigel from a −20 °C freezer and thaw it
overnight in a 4 °C refrigerator.
7. Place Matrigel on ice to avoid premature gelation and dilute the Matrigel stock
solution (typically 9 mg/mL) with cold growth medium to obtain a Matrigel
working solution (2 mg/mL).
8. Prepare 1–4 million cells/mL in 0.75 % alginate containing 0.5 mg/mL Matrigel
by mixing 1 mL of 2–8 million cells/mL, 0.5 mL of 3 % alginate in distilled
water, and 0.5 mL of Matrigel working solution (2 mg/mL) using a 1 mL
pipette. Note: Do not vortex the mixture as mammalian cells are fragile. Mix
well by aspirating and dispensing the solution repeatedly (20 times) with a
1 mL pipette.
9. Place the micropillar chips with dried PLL-CaCl2 spots on the chilling chip
deck at 4–12 °C. Note: Do not leave the micropillar chips too long on the
chilling chip deck unattended as there might be severe water condensation
observed on the chip surface, which will affect spot attachment.
10. Add 100–200 μL/96-well of 1–4 million cells/mL in 0.75 % alginate containing
0.5 mg/mL Matrigel in a 96-well plate. Note: Always use the same kind of the
96-well plates to avoid tip crashing due to Z height difference in 96-wells.
11. Immediately before cell printing, resuspend the cells in 0.75 % alginate con-
taining 0.5 mg/mL Matrigel with a pipette. Note: For uniform cell printing
and preventing solenoid valve and tip clogging, it is extremely important to
resuspend the cells immediately before printing using a 100 μL pipette.
12. Print 60 nL of the cell suspension in 0.75 % alginate containing 0.5 mg/mL
Matrigel on the micropillar chips placed on the chilling chip deck at 4–12 °C
using a work file (Cell_Alginate_6 tip_6 block_60nL).
13. After 1–2 min gelation on the chilling chip deck, sandwich the micropillar chip
with encapsulated cells with the microwell chip containing 950 nL of growth
medium.
14. Rinse tubing, solenoid valves, and ceramic tips with cold distilled water by run-
ning the daily washing program twice. Note: Although Matrigel cannot be
4.5 Sample Printing Protocols 95
gelled in alginate at this low concentration, rinsing with cold water can
ensure to remove residual Matrigel.
15. Place the sandwiched chips with the micropillar chip on the top in a humid cell
incubation chamber or a petri dish with 10 mL water (Fig. 4.8B) and incubate
the chips in a 5 % CO2 incubator at 37 °C for 3D cell culture. Note: Unless
cytotoxicity noticed due to CaCl2, there will be no need to rinse the micropil-
lar chip with growth medium.
16. If necessary, inspect the sandwiched chips under a bright field microscope with
4× or 10× objective lens for spot attachment and uniform cell printing. Note:
Make sure to avoid water evaporation during chip inspection.
17. Change growth medium every 1–2 days by printing fresh growth medium in the
microwell chip, discarding the microwell chip containing old growth medium,
and sandwiching the freshly prepared microwell chip.
1. Coat the micropillar chip with 2 mL of 0.01 % (w/v) PMA-OD in ethanol and
dry it at room temperature at least for 2–3 h.
2. Sonicate 500 μL of 1 % PuraMatrix (from BD Biosciences) in a 1.5 mL
Eppendorf tube for 30 min every time to decrease the viscosity. Note: If air
bubbles are present, centrifuge the PuraMatrix stock solution.
3. Prepare 0.25 % PuraMatrix by mixing 1 % PuraMatrix with sterile deionized
water in a 1:3 (v/v) ratio.
4. Print 60 nL of 0.25 % PuraMatrix on the micropillar chip using a work file
(60 nL PLL_CaCl2) and dry it for at least 4 h. Note: If the micropillar chips
with dried PuraMatrix are not being used within 24 h of the print, store over-
night at 4 °C.
5. For Hep3B cells, prepare RPMI medium containing 10 % FBS, 1 % Pen/Strep
(100 x dilution), and 50 μg/mL gentamycin (1000 × dilution).
6. Print 950 nL of growth medium onto the microwell chip using a work file
(medium_6 tip_6 block) while maintaining the chilling chip deck at 4–12 °C
and incubate it in a humid chamber at 37 °C until use to avoid water evaporation.
Note: Three microwell chips with growth medium are necessary for every
micropillar chip with cells encapsulated in PuraMatrix.
7. Mix PuraMatrix stock with 20 % sucrose in sterile deionized water in a 1:1
(v/v) ratio to generate a 0.5 % PuraMatrix solution in 10 % sucrose.
8. Trypsinize Hep3B cells, mix them with 7 mL growth media, and spin down cell
suspension at 1200 rpm (200 g) for 4 min.
9. Remove the supernatant and resuspend cell pellet with 3 mL of growth media.
10. Take a small aliquot and measure cell density using a cell counter. Note:
Sucrose might be toxic to cells, and interferes cell counting. Thus, do not use
solutions containing sucrose yet.
11. Seed cells in a T75 flask for near future experiments, if necessary.
12. Recentrifuge the cell suspension for the same speed and duration as described
in step 8, completely remove growth media from cell pellet to eliminate salts,
and resuspend the cells with 7 mL of 10 % sucrose in sterile deionized water to
maintain osmolarity. Note: Resuspension of the cells must be done very gently
with the 10 % sucrose solution as sucrose may lyse the cells.
13. Centrifuge the cells at 200 g for 7 min, remove the supernatant, and then resus-
pend the cells in 10 % sucrose at 2× the final desired cell concentration (12 mil-
lion cells/mL). Note: The gelation of PuraMatrix is initiated by salt
concentrations greater than 1 mM in buffers or growth media. Therefore, do
not combine PuraMatrix with salt-containing buffers or growth media until
gelation is desired.
14. Mix the 0.5 % PuraMatrix solution in 10 % sucrose with the 2× cell suspension
in 10 % sucrose in a 1:1 (v/v) ratio very gently for a final concentration of
6 million cells/mL in 0.25 % PuraMatrix and 10 % sucrose. Note: Vigorous
mixing can lyse the cells and release intracellular salts, resulting in premature
gelation of PuraMatrix.
15. Immediately after preparing the mixture freshly, print 60 nL of the cell suspension
in 0.25 % PuraMatrix and 10 % sucrose on the micropillar chip with dried
98 4 Biological Sample Printing
Recombinant viruses have been used to transfer genes to human cells encapsulated
in hydrogels on the micropillar chip. For example, recombinant lentiviruses with
genes for neural stem cell (NSC)-specific biomarkers, including Synapsin1 for neu-
ron differentiation, glial fibrillary acidic protein (GFAP) for astrocyte differentia-
tion, myelin basic protein (MBP) for oligodendrocyte differentiation, and SOX2 for
self-renewal, have been used on the chip platform to monitor real-time NSC differ-
entiation. Recombinant adenoviruses carrying genes for cytochrome P450
(CYP450), including CYP1A2, CYP2C9, CYP2D6, CYP2E1, and CYP3A4, have
been constructed and used on the chip platform for transiently controlled expression
of CYP450 isoforms for metabolism-induced toxicity assays. Protocols for printing
viruses are simple and straightforward, compared to printing cells and enzymes.
Prior to virus printing, it is important to know the number of virus particles in the
solution. The multiplicity of infection (MOI) which is the number of virus particles
per cell has to be measured to precisely control the expression levels of recombinant
viruses used and minimize the basal toxicity due to excessive use of the viruses.
1. Seed HEK293 cells at a density of 10,000 cells in 100 μL Eagle’s minimum
essential medium (EMEM) per well in a 96-well plate.
2. Incubate the 96-well plate in a 5 % CO2 incubator at 37 °C for 24 h. Check to see
if cells are attached to the surface of 96 wells and spread out.
4.5 Sample Printing Protocols 99
Controls
1 2 3 4 5 6 7 8 9 10 11 12
A (10-12)
B (10-11)
C (10-10)
D (10-9)
E (10-8)
F (10-7)
G (10-6)
H (10-5)
Fig. 4.10 Layout of a 96-well plate with different virus dilution in growth medium. The dilution
factors used in the rows are recommended ranges. Sometimes there is need to increase/decrease the
dilution factors accordingly. Growth medium alone (without viruses) is used as a control
3. Dilute viruses with EMEM at 10−5–10−12 fold dilution (Fig. 4.10) and incubate
the 96-well plate in the 5 % CO2 incubator at 37 °C for 7–10 additional days.
4. After 7–10 days of incubation, check each well for cytopathic effect (CPE). This
assay is only valid if the following are true.
1) The control wells (columns 11 and 12) are completely CPE negative.
2) The highest viral concentration (row H1–10) is completely CPE positive.
3) The lowest viral concentration (row A1–10) is completely CPE negative.
5. Sum the fraction of wells in individual row that exhibit CPE.
6. Calculate viral titer by using the following equation:
H
0.8 + n + å fi .
æ pfu ö
Titer ç ÷ = 10 i= A
è mL ø
Fig. 4.11 Layout of the microwell chip containing different MOI of adenoviruses for gene expres-
sion. For example, C1 is no virus condition, C2 is 5 MOI of adenovirus carrying gene for GFP
(Ad-GFP), C3 is 20 MOI of Ad-GFP, C4 is 10 MOI of Ad-GFP plus 10 MOI of Ad-RFP, C5 is 5
MOI of Ad-RFP, and C6 is 20 MOI of Ad-RFP
1. To inspect cells printed on the micropillar chip or in the microwell chip under a
bright-field microscope (EVOS XL Core Imaging System, ThermoFisher
Scientific), turn on the microscope by pressing the on/off switch on the right side
of the microscope and wait until the starting messages disappear and the screen
is ready for inspection (Fig. 4.12).
2. Select an objective lens for chip inspection. There are three different magnifica-
tions of objective lenses (including 4×, 10×, and 20×) installed in the bright-
field microscope. Note: Typically, 4× objective lens is used for rapid inspection
of spot detachment, and 20× objective lens is used for inspecting cell
morphology.
3. Place the chip with cells on the chip deck and adjust focus positions by rotating
the coarse and fine focus adjustment knob for visualizing cells (Fig. 4.13a).
Note: Make sure to avoid water evaporation during chip inspection.
4. Move around the chip with cells to inspect spot detachment and uniform cells
printing in each block. Note: Make sure that the overall number of cells in
Fig. 4.12 Picture of the microscope with the on/off switch, objective lenses, and the focus adjust-
ment knob: (Left) front view and (Right) side view
102 4 Biological Sample Printing
Fig. 4.13 (a) Adjusting focus positions, (b) freezing the image, and (c) saving the cell image on
the chip.
spots printed by different solenoid valves are similar. The block-to-block varia-
tion (i.e., solenoid valve-to-solenoid valve variation) should be less than
20 %. In addition, spot detachment should be below 1 %.
5. Press the “Freeze” button to freeze the cell image before taking a picture
(Fig. 4.13b).
6. Press the “Save” button to store the picture in the USB connected to the micro-
scope (Fig. 4.13c). Note: Do not move the chip until the save message disap-
pears from the screen.
For robust assay development on the chip platform, it is important to measure the
range of errors and chip-to-chip and day-to-day reproducibility. The coefficient of
variation (CV) and the Z’ factor are commonly measured to evaluate error ranges
and robustness of an assay. The CV is defined as the ratio of the standard deviation
(σ) to the average (μ).
s
CV =
m
It is the extent of variability in relation to the average of the signal, thus the
inverse of the signal-to-noise ratio. The range of acceptable CV is typically
below 20 %.
References 103
Instead of using signal to noise (S/N) or signal to background (S/B), the Z’ factor
is the most widely used to measure the robustness of a new assay. The Z’ factor is
defined as the following equation where the averages of maximum/minimum sig-
nals from positive/negative controls are μ+/μ− and the standard deviations of posi-
tive/negative controls are σ+/σ−.
3 (s + + s - )
Z¢ = 1 -
( m+ - m- )
Assays with a Z’ factor between 0.5 and 1.0 are considered excellent. In addition,
assays with a Z’ factor between 0 and 0.5 are considered marginal. The Z’ factor is
used to identify the quality of a new assay prior to testing a large numbers of com-
pounds using high-throughput screening (HTS).
4.8 Summary
In this chapter, we have reviewed the essential methods for printing biological sam-
ples. These methods were optimized for each individual sample, as a suspension of
cells in media will have a different viscosity than cells suspended in each of the vari-
ous hydrogels. Additionally, we have optimized preparation and printing protocols
for hydrogels with various gelation mechanisms, including ion-responsive, pH-
responsive, enzymatically-activated, and thermos-responsive hydrogels. Optimizing
these protocols are essential for ensuring cell viability and reproducing mechanical
and physical interactions observed between cells and ECM in vivo. Additionally,
with successful enzyme, virus, and compound printing protocols established, it is
possible to transiently control gene expression, and predict drug efficacy and toxic-
ity for current and future therapeutics.
References
1. Lee, D. W., Lee, M. Y., Ku, B., Yi, S. H., Ryu, J. H., Jeon, R., et al. (2014). Application of the
DataChip/MetaChip technology for the evaluation of ajoene toxicity in vitro. Archives of
Toxicology, 88(2), 283–290. doi:10.1007/s00204-013-1102-9.
2. Lee, M.-Y., Kumar, R. A., Sukumaran, S. M., Hogg, M. G., Clark, D. S., & Dordick, J. S.
(2008). Three-dimensional cellular microarray for high-throughput toxicology assays.
Proceedings of the National Academy of Sciences, 105(1), 59–63. doi:10.1073/
pnas.0708756105.
3. Lee, M.-Y., Park, C. B., Dordick, J. S., & Clark, D. S. (2005). Metabolizing enzyme toxicology
assay chip (MetaChip) for high-throughput microscale toxicity analyses. Proceedings of the
National Academy of Sciences, 102(4), 983–987. doi:10.1073/pnas.0406755102.
104 4 Biological Sample Printing
4. Kadletz, L., Heiduschka, G., Domayer, J., Schmid, R., Enzenhofer, E., & Thurnher, D. (2015).
Evaluation of spheroid head and neck squamous cell carcinoma cell models in comparison to
monolayer cultures. Oncology Letters, 1281–1286. doi:10.3892/ol.2015.3487.
5. Pawar, S. N., & Edgar, K. J. (2012). Alginate derivatization: A review of chemistry, properties
and applications. Biomaterials, 33(11), 3279–3305. doi:10.1016/j.biomaterials.2012.01.007.
6. Jeon, O., Bouhadir, K. H., Mansour, J. M., & Alsberg, E. (2009). Photocrosslinked alginate
hydrogels with tunable biodegradation rates and mechanical properties. Biomaterials, 30(14),
2724–2734. doi:10.1016/j.biomaterials.2009.01.034.
7. Jeon, O., Powell, C., Ahmed, S. M., & Alsberg, E. (2010). Biodegradable, photocrosslinked
alginate hydrogels with independently tailorable physical properties and cell adhesivity. Tissue
Engineering. Part A, 16(9), 2915–2925. doi:10.1089/ten.tea.2010.0096.
8. Hsiong, S. X., Boontheekul, T., Huebsch, N., & Mooney, D. J. (2009). Cyclic arginine-glycine-
aspartate peptides enhance three-dimensional stem cell osteogenic differentiation. Tissue
Engineering. Part A, 15(2), 263–272. doi:10.1089/ten.tea.2007.0411.
9. Morritt, A. N., Bortolotto, S. K., Dilley, R. J., Han, X., Kompa, A. R., McCombe, D., et al.
(2007). Cardiac tissue engineering in an in vivo vascularized chamber. Circulation, 115(3),
353–360. doi:10.1161/CIRCULATIONAHA.106.657379.
10. Ponce, M. L. (2009). Tube formation: An in vitro matrigel angiogenesis assay. Methods in
Molecular Biology, 467, 183–188.
11. Thiele, J., Ma, Y., Bruekers, S. M. C., Ma, S., & Huck, W. T. S. (2014). 25th anniversary arti-
cle: Designer hydrogels for cell cultures: A materials selection guide. Advanced Materials,
26(1), 125–148. doi:10.1002/adma.201302958.
12. Datar, A., Joshi, P., & Lee, M. Y. (2015). Biocompatible hydrogels for microarray cell printing
and encapsulation. Biosensors, 5(4), 647–663. doi:10.3390/bios5040647.
13. Zhang, Z., He, Q., Deng, W., Chen, Q., Hu, X., Gong, A., et al. (2015). Nasal ectomesenchymal
stem cells: Multi-lineage differentiation and transformation effects on fibrin gels. Biomaterials,
49, 57–67. doi:10.1016/j.biomaterials.2015.01.057.
14. Eyrich, D., Brandl, F., Appel, B., Wiese, H., Maier, G., Wenzel, M., et al. (2007). Long-term
stable fibrin gels for cartilage engineering. Biomaterials, 28(1), 55–65. doi:10.1016/j.
biomaterials.2006.08.027.
15. Luyckx, V., Dolmans, M.-M., Vanacker, J., Scalercio, S. R., Donnez, J., & Amorim, C. A.
(2013). First step in developing a 3D biodegradable fibrin scaffold for an artificial ovary.
Journal of Ovarian Research., 6(1). doi:10.1186/1757-2215-6-83.
16. Huang, Y.-C. Y., Dennis, R. R. G. R. R. G., Larkin, L., & Baar, K. (2005). Rapid formation of
functional muscle in vitro using fibrin gels. Journal of Applied Physiology, 98(2), 706–713.
doi:10.1152/japplphysiol.00273.2004.
Chapter 5
High-Content Cell Staining
Contents
5.1 Introduction...................................................................................................................... 105
5.1.1 Fluorescent Dyes.................................................................................................. 106
5.1.2 Immunofluorescence (IF) Assays with Antibodies.............................................. 107
5.1.3 Fluorescent Proteins............................................................................................. 108
5.2 Materials.......................................................................................................................... 109
5.2.1 Reagents for Fluorescence Staining..................................................................... 109
5.2.2 Reagents for Immunofluorescence Staining........................................................ 109
5.2.3 Devices for Cell Staining..................................................................................... 110
5.2.4 Preparation of Dye Stock Solutions in DMSO.................................................... 111
5.3 Protocols.......................................................................................................................... 112
5.3.1 Staining Cells with Fluorescent Dyes.................................................................. 112
5.3.1.1 Preparation of a Saline Solution.......................................................... 112
5.3.1.2 Staining Cells on the Micropillar/Microwell Chip Platform
with Fluorescent Dyes......................................................................... 112
5.3.2 Staining Cells on the Chip Platform with Fluorophore-Labeled Antibodies....... 116
5.3.2.1 Cell Fixation........................................................................................ 117
5.3.2.2 Permeabilization of Cell Membranes.................................................. 117
5.3.2.3 Blocking of Nonspecific Binding and Incubation with Primary/
Secondary Antibodies for Fluorescence Labeling and Detection....... 118
5.3.3 Measuring the Expression Levels of Drug Metabolizing Enzymes on a Chip.... 119
5.3.3.1 Tyramide Signal Amplification Kit (Life Technologies).................... 121
5.4 Summary.......................................................................................................................... 122
References................................................................................................................................. 122
5.1 Introduction
analysis algorithms [2]. In particular, high-content imaging (HCI) has gained popu-
larity for systematic and accurate evaluation of drug candidates [3, 4] because of its
capability to assess specific signals, including changes in the nucleus, organelle
structure, protein translocation, oxidative stress, apoptosis/necrosis, mitochondrial
impairment, calcium homeostasis, morphology, and phenotype profiling as readouts
[2, 4]. Three different types of cell labeling are typically used for HCI assays, which
include fluorescent dyes for direct cell staining, antibodies for immunofluorescent
labeling, and genetically expressed fluorescent proteins such as green fluorescent
protein (GFP) [1, 5]. This chapter summarizes basic cell staining protocols with
various fluorescent dyes and antibodies for HCI assays on the micropillar/microwell
chip platform.
As very few fluorescent dyes can stain proteins expressed within cells or secreted by
cells specifically, immunofluorescent labeling with antibodies has been developed
to resolve this issue and visualize the distribution of target proteins in biological
samples. Target proteins can interact with antibodies, which belong to a family of
globular proteins called immunoglobulins (Igs) (Fig. 5.1). Ig has four polypeptide
chains divided into two heavy (~50 kD, H) chains and two light (~23 kD, L) chains,
one of which has a specific binding site for the epitope of an antigen (i.e., target
protein). These specific binding of an antibody to an antigen is the major principle
of IF assays [17].
Two types of antibodies such as monoclonal (mAbs) and polyclonal antibodies
(pAbs) are commonly used for IF assays. The major difference is that a mAb
recognizes one epitope on an antigen whereas a pAb recognizes multiple epitopes on
any one antigen. Thus, a mAb is used to detect a specific target protein which leads
to reproducible cell staining with significantly lower background fluorescence. On
the other hand, a pAb is used to detect a target protein with low expression levels and
amplify fluorescent signals due to multiple binding sites. The difference in their
properties are summarized in Table 5.1 [17].
The major advantages of IF staining are their wide applications with sectioned
tissues, primary cells, and cultured cell lines as well as their capability to detect
proteins, glycans, and small biological molecules. The IF-labeled samples can be
Table 5.1 Properties of monoclonal antibody (mAb) and polyclonal antibody (pAb)
mAb pAb
• High specificity, detects only one • High affinity, binds multiple epitopes on any one
epitope on an antigen antigen and amplifying signals for robust detection
for immunoprecipitation (IP) and chromatin
immunoprecipitation (ChIP)
• Reduces background staining and • High background and potentially generates
highly reproducible between inaccurate results
experiments
• Used as primary antibody • Used as secondary antibody
• Expensive and takes a long time • Inexpensive and relatively easy to produce
for hybridoma production
• Too susceptible to the loss of • High batch-to-batch variability
epitope after chemical fixation
Cells expressing fluorescent proteins are used in HCI assays to investigate various
cellular processes, such as protein trafficking, gene activation, protein-protein
interactions, organelle condition, and cellular development/differentiation [5].
Fluorescent proteins in mammalian cells are expressed by transferring genes encod-
ing fluorescent proteins into the cells using plasmids, viral vectors, or other means.
The processes known as transfection or transduction can be either transient or
permanent. In transient transfection/transduction, the gene introduced into the cell
does not integrate into the chromosomes, therefore expressing the protein for a short
period of time. On the other hand, the gene can be permanently incorporated into
the genome to create transformed cell lines. Selection of fluorescent proteins for
HCI generally depends on several properties, including efficiency of protein expres-
sion, toxicity of the protein in the cells, brightness, photostability, and insensitivity
5.2 Materials 109
5.2 Materials
Anti-mouse IgG (H+L), highly cross-adsorbed, Alexa Fluore® 488 antibody produced in donkey
• Hoechst 33342
–– Dissolve 20 mg of Hoechst 33342 in 3.25 mL DMSO to prepare a stock solution
of 10 mM Hoechst 33342.
–– Store aliquots of 10 mM Hoechst 33342 in a −20 °C freezer protected from
light until use.
–– To prepare a working solution of 25 μM Hoechst 33342, add 20 μL of the stock
solution in 8 mL of Dulbecco’s phosphate buffered saline (DPBS). Note: If
possible, do not reuse the stock solution of the dye for reproducible results.
• Tetramethyl rhodamine methyl ester (TMRM)
–– Dissolve 25 mg TMRM in 1 mL DMSO and vortex it for 1 min to prepare a
stock solution of 50 mM TMRM.
–– Store aliquots of 50 mM TMRM in a −20 °C freezer protected from light
until use.
–– Dilute the stock solution 100-fold to get a working stock concentration of
0.5 mM. Store aliquots of 0.5 mM TMRM in a −20 °C freezer, and do not
reuse this stock solution for reproducible results.
–– To prepare a working solution of 0.5 μM TMRM, add 8 μL stock in 8 mL
DPBS.
• Fluo-4 acetoxy methyl ester (Fluo-4 AM)
–– Dissolve 50 μg of Fluo-4 AM in 23 μL DMSO to prepare a stock solution of
2 mM Fluo-4 AM. Note: Fluo-4 AM may require addition of 20 % (w/v)
Pluronic F-127 to enhance solubility.
–– Store aliquots of 2 mM Fluo-4 AM in a −20 °C freezer protected from light
until use.
–– To prepare a working solution of 5 μM Fluo-4 AM, add 10 μL stock in 2 mL
DPBS.
• Monochlorobimane (mBCl)
–– Dissolve 25 mg of mBCl in 550 μL DMSO to prepare a stock solution of
200 mM mBCl.
–– Store aliquots of 200 mM mBCl in a −20 °C freezer protected from light until use.
–– To prepare a working solution of 100 μM mBCl, add 4 μL stock in 8 mL
DPBS.
• Calcein AM
–– Calcein AM comes in a liquid form in the live/dead viability/cytotoxicity kit
at a concentration of 4 mM without the need for adding solvent.
–– Store aliquots of calcein AM in a −20 °C freezer protected from light until use.
–– To prepare a working solution of 1 μM calcein AM, add 2 μL stock in 8 mL
DPBS.
112 5 High-Content Cell Staining
• Ethidium homodimer-1
–– Ethidium homodimer-1 comes in a liquid form in the live/dead viability/cyto-
toxicity kit at a concentration of 2 mM without the need for adding solvent.
–– Store aliquots of ethidium homodimer-1 in a −20 °C freezer protected from
light until use.
–– To prepare a working solution of 1 μM ethidium homodimer-1, add 4 μL
stock in 8 mL DPBS.
• Propidium iodide
–– Propidium iodide comes in a liquid form at a concentration of 1.5 mM with-
out the need for adding solvent.
–– Store aliquots of 1.5 mM propidium iodide in a 4 °C refrigerator protected
from light until use. Note: Propidium iodide is stable for at least 6 months
when stored at 4 °C.
–– To prepare a working solution of 5 μM propidium iodide, add 27 μL stock in
8 mL DPBS.
• YO-PRO-1
–– YO-PRO-1 comes in liquid form at a concentration of 1 mM without the need
for adding solvent.
–– Store aliquots of 1 mM YO-PRO-1 in a −20 °C freezer protected from light
until use.
–– To prepare a working solution of 10 μM YO-PRO-1, add 80 μL stock in 8 mL
DPBS.
5.3 Protocols
5.3.1.2 S
taining Cells on the Micropillar/Microwell Chip Platform
with Fluorescent Dyes
1. Prepare working solutions of desired fluorescent dyes in the saline solution prior
to initial washing of the chip, and keep it protected from light (Table 5.3).
5.3 Protocols
Fig. 5.2 Pictures of (a) an empty deep-well plate, (b) a deep-well plate with a saline solution and
(c) a deep-well plate with the micropillar chips placed on the saline solution for cell rinsing.
2. Using 10 mL serological pipette, aspirate the saline solution and dispense 5.5 mL
in each well of the deep-well staining plate (Fig. 5.2B). Note: Make sure to avoid
dust particles while washing and staining. To remove dust and tiny particles
from deep-well and shallow-well plates, blow air on the wells of both deep-well
and shallow-well staining plates before starting the cell staining process.
3. Place the micropillar chip facing down on top of the deep well with 5.5 mL of
the saline solution so that cells on the micropillars can be immersed in the solu-
tion, and rinse the chip twice for 5 min each (Fig. 5.2C). Note: The cells on the
chip are immersed in the saline solution to remove growth media and fetal
bovine serum (FBS), which may interfere cell staining. Make sure that there
are no bubbles trapped underneath the chip.
4. Add 2 mL of a desired fluorescent dye solution in each well of a shallow-well
plate (Fig. 5.4B). Note: Spread the dye solution with a pipette tip evenly on the
surface of the shallow well. As the surface of the shallow well is relatively
hydrophobic, it may not be spread well enough for uniform cell staining.
5. Once the chip is rinsed on the deep-well plate, drain an excess saline solution
from the chip by tilting the chip at an angle of 45° and remove the remaining
saline solution from the side of the chip with a paper towel (Fig. 5.3). Note:
Do not dry the cell spots on the micropillar chip as spot drying will cause
cell death.
6. Immediately after draining the excess saline solution, place the micropillar chip
facing down on top of the shallow well with 2 mL of the desired fluorescent dye
solution and incubate for 60 min for cell staining (Fig. 5.4C). Note: The staining
time can be varied depending on the fluorescent dye used and the morphology
of cells stained.
5.3 Protocols 115
Fig. 5.3 Removing the excess saline solution from the side of the micropillar chip with a paper
towel by tilting the chip at a 45° angle
Fig. 5.4 Pictures of (a) an empty shallow-well plate, (b) a shallow-well plate with a fluorescent
dye solution and (c) a shallow-well plate with the micropillar chips placed on the fluorescent dye
solution for cell staining
7. During the period of cell staining, cover the entire shallow-well plate with the
chips with aluminum foil. Note: The majority of fluorescent dyes are very sensi-
tive to light illumination. To avoid photobleaching of fluorescent dyes during
staining, make sure to prevent the chips from light exposure.
8. After cell staining, remove an excess dye solution by rinsing the micropillar
chip with stained cells twice for 10 min each in the deep-well plate with 5.5 mL
of the saline solution. The micropillar chips should be protected from light dur-
ing the rinsing steps too.
9. After rinsing, drain and remove the excess saline solution from the chip as
described in Step 5.
116 5 High-Content Cell Staining
Fig. 5.5 Representative images of Hep3B cells cultured in 3D (60 nL) on the micropillar chip and
stained with (b) calcein AM for cell viability, (b) TMRM for mitochondrial membrane potential,
(c) Hoechst 33342 for nucleus morphology and cell count, (d) mBCl for glutathione level, (e)
propidium iodide for cell viability and necrosis, and (f) Fluo-4 AM for intracellular calcium level
10. Dry the stained chips completely in the dark for at least 2 h. Once the micropillar
chips are completely dried, they are ready for scanning. As an example, images
of 3D-cultured cells stained with various fluorescent dyes are presented in Fig.
5.5. Note: Do not scan a wet micropillar chip as dripping water with salts can
damage objective lenses and filter sets in the S+ Scanner. Always completely
dry the micropillar chip or seal the microwell chip with wet samples with a gas-
permeable sealing membrane (Sigma-Aldrich) before scanning.
1. Rinse the cells by immersing the micropillar chip containing cells twice in
5.5 mL of 1× sterilized PBS (or 1× TBS) in the deep-well plate for 5 min each.
2. Incubate the chip in the shallow-well plate with 2 mL of a fixation reagent for
10–15 min. Note: Refer Table 5.4 for various fixation reagents and their
methods.
3. Rinse the cells three times in 5.5 mL of 1× sterilized PBS (or 1× TBS) in the
deep-well plate for 3 min each.
4. Incubate cells with permeabilization reagents for 10 min. Note: Refer Table 5.5
for various permeabilization reagents and their methods.
5. After permeabilization, rinse cells with 1× PBS or 1× TBS.
6. After cell fixation and permeabilization, incubate the cells in 1× PBS (or 1×
TBS) containing 3 % BSA (or other blocking reagents) for 1 h to prevent non-
specific binding of antibodies. Note: Blocking reagents may vary depending on
the application. Refer to Table 5.6 for various blocking reagents and their
methods of use.
5.3 Protocols 119
Table 5.7 Examples of primary antibodies for detecting cytochrome P450 isoforms
Name Target protein Origin Dilution factor Company
Rabbit polyclonal anti-CYP1A1 CYP1A1 Rabbit 1:1000 Abcam
Rabbit polyclonal anti-CYP3A4 CYP3A4 Rabbit 1:100 Thermo Fisher
Scientific
Rabbit polyclonal anti-CYP2C9 CYP2C9 Rabbit 1:100 Thermo Fisher
Scientific
by touching the bottom of the petri dish when lifting the floating chip up with
a pair of tweezers. Do not shake in order to avoid spot detachment.
3. Fix and permeate the cells by immersing the chip in 25 mL of 1× TBS containing
3.7 % formaldehyde for 20 min at room temperature and then transferring the
chip in 25 mL of 1× TBS containing 0.15 % Triton X-100 for 10 min. In case the
chip made of glass is used, fix and permeate the cells by immersing the glass chip
in 25 mL of methanol and acetone (1:1, v/v) for 20 min at −20 °C.
4. Rinse the cells by immersing the chip twice in 25 mL of 1× TBS containing
0.15 TBS for 5 min each.
5. Dry the chip under weak nitrogen gas (or air) stream.
6. Incubate the chip overnight in 25 mL of a blocking buffer in TBS (SuperBlock
from Fisher Scientific) at 4 °C.
7. Wash the chip three times in 25 mL of TBS-T for 5 min each. Note: Tween 20 in
TBS-T is necessary to reduce hydrophobic interactions between protein and
surface.
8. Label the cells with a primary antibody (e.g., mouse anti-human IgG from Life
Technologies) by diluting the primary antibody (typically 1:100–1:500, v/v, see
primary antibody dilutions used in Table 5.7 as a reference) in TBS-T containing
1 % (w/v) bovine serum albumin (BSA), applying 2 mL of the primary antibody
solution to the shallow-well plate, and then incubating the chip stamped onto the
staining plate overnight at 4 °C.
9. Rinse the cells by immersing the chip three times in 25 mL of TBS-T for 15 min
each with gentle shaking.
10. Label the cells with a secondary antibody (e.g., HRP-conjugated goat anti-
mouse IgG from Life Technologies T20912) by diluting the secondary anti-
body (typically 1:500–1:1000, v/v) in TBS-T containing 1 % BSA, applying
2 mL of the secondary antibody solution to the shallow-well plate, and then
incubating the chip stamped onto the staining plate for 3 h at room
temperature.
11. Rinse the cells by immersing the chip three times in 25 mL of TBS-T for 15 min
each with gentle shaking.
12. Label the cells with Alexa Fluor 488 tyramide (Tyramide signal amplification
kit, Life Technologies T20912) by diluting the tyramide stock (1:200, v/v) in
amplification buffer containing 0.0015 % H2O2 just prior to labeling, applying
5.3 Protocols 121
Contents
• Labeled tyramide (Component A), one vial
• Dimethylsulfoxide (DMSO; Component B), 200 μL
• HRP-conjugated secondary antibody (Component C), 100 μg
• Blocking reagent (Component D), 3 g
• Amplification buffer (Component E), 25 mL (containing 0.02 % thimerosal)
• Hydrogen peroxide (H2O2; Component F), 200 μL of a 30 % stabilized solution
Upon receipt and prior to use, the kit should be stored at −20 °C, desiccated and
protected from light.
Preparation of Solutions
1. Prepare tyramide stock solution by dissolving the solid material provided
(Component A) in 150 μL of DMSO (Component B). Invert the vial several
times to dissolve any tyramide coating the sides of the vial. Store unused por-
tions of this stock solution in small aliquots at −20 °C, desiccated and protected
from light.
2. Prepare the HRP-conjugated antibody stock solution by reconstituting the mate-
rial provided (Component C) in 200 μL of PBS. This solution may be stored at
2–6 °C for up to 3 months if required. Optionally, add 0.02 % thimerosal as a
preservative. Note: Sodium azide must not be used for this purpose.
3. Prepare a 1 % (10 mg/mL) solution of blocking reagent in PBS. Note: We recom-
mend preparing only as much as is needed for immediate use. However, unused
solution can be stored frozen at −20 °C for 1 month if necessary.
4. Prepare amplification buffer/0.0015 % H2O2 by adding 30 % hydrogen peroxide
(Component F) to amplification buffer (Component E) to obtain a final concen-
tration of 0.0015 % H2O2. For example, add 5 μL of 30 % H2O2 to 995 μL of
amplification buffer and then add 20 μL of this intermediate dilution (0.15 %
H2O2) to a further 1980 μL of amplification buffer.
122 5 High-Content Cell Staining
5.4 Summary
References
1. Charvin, G., Oikonomou, C., & Cross, F. R. (2010). Labels and probes for live cell imaging:
Overview and selection guide. Methods in Molecular Biology, 591(2), 229–242.
doi:10.1007/978-1-60761-404-3_14.
2. van Vliet, E., Danesian, M., Beilmann, M., Davies, A., Fava, E., Fleck, R., et al. (2014).
Current approaches and future role of high content imaging in safety sciences and drug discovery.
ALTEX, 31(4), 479–493. doi:10.14573/altex.1405271.
3. Buchser, W., Collins, M., Garyantes, T., Guha, R., Haney, S., Lemmon, V., et al. (2012). Assay
development guidelines for image-based high content screening, high content analysis and
high content imaging. In G. S. Sittampalam, N. P. Coussens, H. Nelson, M. Arkin, D. Auld,
C. Austin, et al. (Eds.), Assay guidance manual (pp. 1–69). Bethesda, MD: Eli Lilly &
Company and the National Center for Advancing Translational Sciences.
4. Zanella, F., Lorens, J. B., & Link, W. (2010). High content screening: Seeing is believing.
Trends in Biotechnology, 28(5), 237–245. doi:10.1016/j.tibtech.2010.02.005.
5. Valeur, B., & Berberan-Santos, M. N. (2012). Autofluorescence and fluorescence labeling in
biology and medicine. Molecular Fluorescence: Principles and Applications, 479–505.
6. Demchenko, A. P. (2008). Introduction to fluorescence sensing. Berlin: Springer.
doi:10.1007/978-1-4020-9003-5.
7. Virus, E. D., Sobolevsky, T. G., & Rodchenkov, G. M. (2008). Introduction to fluorescent
techniques. Journal of Mass Spectrometry, 43(7), 949–957. doi:10.1002/jms.1447.
8. Towne, D. L., Nicholl, E. E., Comess, K. M., Galasinski, S. C., Hajduk, P. J., & Abraham, V. C.
(2012). Development of a high-content screening assay panel to accelerate mechanism of
action studies for oncology research. Journal of Biomolecular Screening, 17(8), 1005–1017.
doi:10.1177/1087057112450050.
9. Martin, H. L., Adams, M., Higgins, J., Bond, J., Morrison, E. E., Bell, S. M., et al. (2014).
High-content, high-throughput screening for the identification of cytotoxic compounds based
on cell morphology and cell proliferation markers. PLoS ONE, 9(2), 1–8. doi:10.1371/journal.
pone.0088338.
10. Alonso-Padilla, J., Cotillo, I., Presa, J. L., Cantizani, J., Peña, I., Bardera, A. I., et al. (2015).
Automated high-content assay for compounds selectively toxic to trypanosoma cruzi in a myo-
blastic cell line. PLoS Neglected Tropical Diseases, 9, 1–17. doi:10.1371/journal.pntd.0003493.
11. Sirenko, O., Hesley, J., Rusyn, I., & Cromwell, E. F. (2014). High-content assays for hepato-
toxicity using induced pluripotent stem cell-derived cells. Assay and Drug Development
Technologies, 12(1), 43–54. doi:10.1089/adt.2013.520.
12. Håkanson, M., Cukierman, E., & Charnley, M. (2014). Miniaturized pre-clinical cancer mod-
els as research and diagnostic tools. Advanced Drug Delivery Reviews, 69–70, 52–66.
doi:10.1016/j.addr.2013.11.010.
References 123
13. Mioulane, M., Foldes, G., Ali, N. N., Schneider, M. D., & Harding, S. E. (2012). Development
of high content imaging methods for cell death detection in human pluripotent stem cell-
derived cardiomyocytes. Journal of Cardiovascular Translational Research, 5, 593–604.
doi:10.1007/s12265-012-9396-1.
14. Fujisawa, S., Romin, Y., Barlas, A., Petrovic, L. M., Turkekul, M., Fan, N., et al. (2014).
Evaluation of YO-PRO-1 as an early marker of apoptosis following radiofrequency ablation of
colon cancer liver metastases. Cytotechnology, 66(2), 259–273. doi:10.1007/
s10616-013-9565-3.
15. Donato, M. T., Tolosa, L., Jiménez, N., Castell, J. V., & Gómez-Lechón, M. J. (2012). High-
content imaging technology for the evaluation of drug-induced steatosis using a multiparamet-
ric cell-based assay. Journal of Biomolecular Screening, 17(3), 394–400.
doi:10.1177/1087057111427586.
16. Lakowicz, J. R. (2009). Fluorophores. Princ Fluoresc Spectrosc, 954. doi:10.1002/
smll.201090041.
17. Lipman, N. S., Jackson, L. R., Weis-Garcia, F., & Trudel, L. J. (2005). Monoclonal versus
polyclonal antibodies: Distinguishing characteristics, applications, and information resources.
ILAR Journal, 46(3), 258–268. doi:10.1093/ilar.46.3.258.
18. Aubin, J. E. (1979). Autofluorescence of viable cultured mammalian cells. The Journal of
Histochemistry and Cytochemistry, 27(1), 36–43.
19. Ntziachristos, V. (2010). Going deeper than microscopy: the optical imaging frontier in biol-
ogy. Nature Methods, 7(8), 603–614. doi:10.1038/nmeth.1483.
20. Lee, K., Choi, S., Yang, C., Wu, H.-C., & Yu, J. 2013. Autofluorescence generation and elimi-
nation: A lesson from glutaraldehyde. Chemical Communications (Cambridge, England),
49(29):3028–3030. doi:10.1039/c3cc40799c.
21. Shaner, N. C., Steinbach, P. A., & Tsien, R. Y. (2005). Aguide to choosing fluorescent proteins.
Nature Methods, 2(12), 905–909. doi:10.1038/nmeth819.
22. Manning, C. F., Bundros, A. M., & Trimmer, J. S. (2012). Benefits and pitfalls of secondary
antibodies: Why choosing the right secondary is of primary importance. PLoS ONE, 7(6).
doi:10.1371/journal.pone.0038313.
23. Kaku, T., Ekem, J. K., Lindayen, C., Bailey, D. J., Van Nostrand, A. W., & Farber, E. (1983).
Comparison of formalin- and acetone-fixation for immunohistochemical detection of
carcinoembryonic antigen (CEA) and keratin. American Journal of Clinical Pathology, 80(6),
806–815.
24. Levitt, D., & King, M. (1987). Methanol fixation permits flow cytometric analysis of immuno-
fluorescent stained intracellular antigens. Journal of Immunological Methods, 96(2), 233–237.
doi:10.1016/0022-1759(87)90319-X.
25. Kosaka, T., Nagatsu, I., JY, W., & Hama, K. (1986). Use of high concentrations of glutaralde-
hyde for immunocytochemistry of transmitter-synthesizing enzymes in the central nervous
system. Neuroscience, 18(4), 975–990. doi:10.1016/0306-4522(86)90112-0.
26. Pollice, A. A., McCoy, J. P., Shackney, S. E., Smith, C. A., Agarwal, J., Burholt, D. R., et al.
(1992). Sequential paraformaldehyde and methanol fixation for simultaneous flow cytometric
analysis of DNA, cell surface proteins, and intracellular proteins. Cytometry, 13(4), 432–444.
doi:10.1002/cyto.990130414.
27. Grizzle, W. E. (2009). Models of fixation and tissue processing. Biotechnic & Histochemistry,
84(5), 185–193. doi:10.3109/10520290903039052.Models.
28. Miething, F., Hering, S., Hanschke, B., & Dressler, J. (2006). Effect of fixation to the degrada-
tion of nuclear and mitochondrial DNA in different tissues. The Journal of Histochemistry and
Cytochemistry, 54(3), 371–374. doi:10.1369/jhc.5B6726.2005.
29. Jamur MC, Oliver C. (2010). Permeabilization of cell membranes. In Oliver C, & Jamur MC
(Eds.), Methods and Protocols (Vol. 588, pp. 63–66). New York: Humana Press.
doi:10.1007/978-1-59745-324-0.
30. Amidzadeh, Z., Behzad Behbahani, A., Erfani, N., Sharifzadeh, S., Ranjbaran, R., Moezi, L.,
et al. (2014). Assessment of different permeabilization methods of minimizing damage to the
124 5 High-Content Cell Staining
adherent cells for detection of intracellular RNA by flow cytometry. Avicenna Journal of
Medical Biotechnology, 6(1), 38–46.
31. Ohsaki, Y., Maeda, T., & Fujimoto, T. (2005). Fixation and permeabilization protocol is criti-
cal for the immunolabeling of lipid droplet proteins. Histochemistry and Cell Biology, 124(5),
445–452. doi:10.1007/s00418-005-0061-5.
32. Misra, D. P., Chaurasia, S., & Misra, R. (2016). Increased circulating Th17 Cells, Serum
IL-17A, and IL-23 in Takayasu Arteritis. Autoimmune Diseases, 2016, 7841718.
doi:10.1155/2016/7841718.
33. Kopen, G. C., Prockop, D. J., & Phinney, D. G. (1999). Marrow stromal cells migrate through-
out forebrain and cerebellum, and they differentiate into astrocytes after injection into neonatal
mouse brains. Proceedings of the National Academy of Sciences of the United States of
America, 96(19), 10711–10716. doi:10.1073/pnas.96.19.10711.
34. Italiano, J. E., Richardson, J. L., Patel-Hett, S., Battinelli, E., Zaslavsky, A., Short, S., et al.
(2008). Angiogenesis is regulated by a novel mechanism: Pro- and antiangiogenic proteins are
organized into separate platelet α granules and differentially released. Blood, 111(3), 1227–
1233. doi:10.1182/blood-2007-09-113837.
35. Kwong, K. F., Schuessler, R. B., Green, K. G., Laing, J. G., Beyer, E. C., Bioneau, J. P., et al.
(1998). Differential expression of gap junction proteins in the canine sinus node. Circulation
Research, 82, 604–612.
36. Bianchi, L., Shen, Z., Dennis, A. T., Priori, S. G., Napolitano, C., Ronchetti, E., et al. (1999).
Cellular dysfunction of LQT5-minK mutants: Abnormalities of I(Ks), I(Kr) and trafficking in
long QT syndrome. Human Molecular Genetics, 8(8), 1499–1507. doi:10.1093/hmg/8.8.1499.
37. Thompson, K., Trowern, A., & Fowell, A. (1998). Primary rat and mouse hepatic stellate cells
express the macrophage inhibitor cytokine interleukin-10 during the course of activation
in vitro. Hepatology, 28(6), 1518–1524. doi:10.1002/hep.510280611.
Chapter 6
3D-Cultured Cell Image Acquisition
Contents
6.1 Introduction...................................................................................................................... 125
6.2 Materials.......................................................................................................................... 127
6.3 Protocols.......................................................................................................................... 127
6.3.1 Daily Operational Procedures.............................................................................. 128
6.3.2 Parameter Setting Procedures.............................................................................. 135
6.3.2.1 Setting the Position of Filters and the Distance of Each Step
for Autofocus...................................................................................... 135
6.3.2.2 Setting XYZ Coordinates for Different Chips
and Objective Lenses.......................................................................... 137
6.4 Summary.......................................................................................................................... 139
6.5 Appendix.......................................................................................................................... 139
References................................................................................................................................. 141
6.1 Introduction
The basic components of any imaging device consist of light sources, detectors,
objective lenses, and optical filters among which light sources and detectors play
critical role in determining fate of any HCI system. Light sources can comprise of
lamps, lasers, and light-emitting diodes (LEDs) with each of them having their own
benefits and limitations. For example, lamps provide a broad excitation source from
UV to IR, but should be replaced frequently due to shorter lifetime and may need
realignment for optimal excitation. Lasers, on the other hand, have longer lifetime,
but may not offer optimal excitation spectra for certain target fluorophores such as
blue fluorescent dyes. LEDs offer both long lifetime and optimal excitation, but are
expensive. Most of currently available image acquisition systems use xenon or mer-
cury lamps as a light source due to economic feasibility and a broad range of spec-
tra. Detectors are another important component of HCI devices with two types of
detectors commonly used. These detectors are charge-coupled devices (CCDs) and
photomultiplier tubes (PMTs). CCD-based detectors are widely used in HCI devices
as it offers excellent acquisition speed (100 frames per second), large dynamic
ranges (>20,000:1), broad spectral sensitivity (400–900 nm and higher), and high
resolution (>2000 × 2000 pixels), but with reduced sensitivity to low intensity light.
On the other hand, PMT-based detectors offer high sensitivity towards low intensity
light, but with compromise in throughput. Multiple PMTs are used in parallel to
acquire images through various fluorescent channels [1–3].
Conventional HCI devices can be divided broadly into two categories; wide-field
imaging devices and confocal imaging devices. Wide-field imaging devices are
basically inverted microscopes which offer an excellent image acquisition speed.
Some examples of commercially available CCD-based wide-field imaging devices
are ArrayScan XTI HCS Reader manufactured by Thermo Fisher, IN Cell Analyzer
2000 by GE Healthcare, ImageXpress Micro HCS system by Molecular Devices,
Operetta by Perkin Elmer, WiSCAN by IDEA Bio-Medical, and S+ Scanner by
SEMCO. These commercially available wide-field imaging devices commonly use
lamps and LEDs as their light source. For example, ArrayScan XTI HCS Reader
uses a metal halide lamp and LED, and IN Cell Analyzer 2000 uses a metal halide
lamp. Similarly, ImageXpress Micro HCS system uses a xenon lamp, and Operetta
uses a xenon lamp and LED. In addition, S+ scanner uses a mercury lamp, and
WiSCAN uses a mercury lamp and LED. Although different types of light sources
are used by various wide-field imaging devices, the difference is very little in terms
of optical performance. The difference between the various wide-field imagers
available arise generally from the numerical aperture of the objectives and the qual-
ity of light path [1, 4, 5].
Confocal imaging devices use light barrier with a fixed or adjustable pinhole to
eliminate light in front or behind the focus plane of an objective. Confocal devices
offer high resolution and improved contrast compared to wide-field devices, but it
also causes reduction in light signal and only scans single points of a specimen at
any given moment. Although these issues have been overcome by using technolo-
gies like laser scanning confocal microscopy (LSCM) and Nipkow spinning disk,
the image acquisition speed is compromised, and there is always a chance of photo-
6.3 Protocols 127
bleaching [1, 4, 6, 7]. Confocal imaging devices are expensive compared to wide-
field imaging devices and most of its applications are towards imaging small
intracellular structures, small cells, complex 3D structures, etc. IN Cell Analyzer
6000 by GE Healthcare is a commonly used confocal microscope, which uses
LSCM with varying apertures. It consists of four laser lines and a LED for transmit-
ted light and a scientific complementary metal oxide semiconductor (sCMOS)-
based detector. ImageXpress ULTRA developed by Molecular devices is another
confocal imager consisting of four lasers and four PMTs that can be operated seri-
ally or in parallel [1, 8, 9]. Extensive information on various HCI devices and their
components can be found in Assay Guidance Manual published by Eli Lilly &
Company and the National Center for Advancing Translational Sciences (http://
www.ncbi.nlm.nih.gov/books/NBK53196/) [1].
While other HCI devices are highly sophisticated and require trained personnel
to operate the machine, the S+ Scanner offers researchers/users the ability to acquire
high quality fluorescent images in a high-throughput manner without the need of
complex training. This chapter aims to provide simple procedures for operating S+
Scanner to rapidly acquire 3D-cultured cell images from the micropillar/microwell
chip platform.
6.2 Materials
6.3 Protocols
Chip-loading deck
Light source
Detector
Objective lens
Filter set
objective lens and stored in the dark to prevent photobleaching. In addition, when
acquiring the image from microwell chips containing liquid samples it should be
covered with a gas-permeable sealing membrane to prevent spilling of the samples
and to minimize immediate drying of the spots.
1. Turn on the power source (by using the blue switch on the step-up transformer).
2. Turn on the S+ Scanner using the green power switch and then press the ‘Reset’
button (Fig. 6.2). Note: Pushing the ‘Reset’ button resets XYZ coordinates of
the chip-loading deck. This step is essential to avoid malfunctioning of the S+
Scanner. Do not skip this step!
3. Turn on the light source and the computer. Use the black switch on the back to
turn on the light source and the ‘ON/OFF’ button in the front to turn on the lamp
(Fig. 6.3). Note: After the lamp is turned ON, do not turn OFF for 2 min to
avoid damage to the lamp. In addition, do not turn ON the lamp within 10 min
of turning OFF. The operation of switches is disabled for 5 min after the lamp
is turned OFF. The lamp is connected to the light source in S+ Scanner via a
fiber optic cable, which needs to be placed properly to avoid damage to the fiber
optic cable.
4. Open the scanner program by double clicking the shortcut of S+ Scanner located
on the desktop (Fig. 6.4).
5. The user interface of the scanner program will be opened (Fig. 6.5).
6. Make sure that the sliding door on top of the scanner is closed (Fig. 6.2).
6.3 Protocols 129
Sliding door/
Chip-loadingdeck
Reset button (Blue)
Emergency button (Red)
Power switch
Fig. 6.2 The picture of S+ Scanner showing the power switch on the back, reset and emergency
buttons on the side, and sliding door/chip-loading deck on the top
Fig. 6.3 Pictures of the light source: (a) The front side with ON/OFF and Reset buttons and a LED
indicator displaying the burner status and (b) The back side with the main switch and the fiber
optic cable
7. In the ‘Home’ window, click the ‘Home’ icon in the ‘XYZ Move’ tab to activate
the ‘Chip loading’ and ‘Live’ icon (Fig. 6.6).
8. Click ‘Chip loading’ icon in the ‘XYZ Move’ tab to bring the chip-loading deck
in position for loading the micropillar chip zig (Fig. 6.6).
Fig. 6.4 The shortcut of S+ Scanner program
Fig. 6.5 The user interface of S+ Scanner program showing ‘Home’ window highlighted in a red
box
Fig. 6.6 ‘Home’ window in the user interface of the scanner program
Fig. 6.7 Preparation of micropillar chips before loading on the chip-loading deck for image acqui-
sition. (a) Placing a metal frame on the micropillar chip at an angle of 45°, (b) Inserting all the
micropillars into the holes on the metal frame, (c) Placing the micropillar chip with the metal frame
on the magnetic micropillar chip zig. The circle highlighted in yellow indicates the bottom of the
zig aligned with the bottom of the micropillar chip.
6.3 Protocols 131
9. In order to place micropillar chips on the micropillar chip zig and load it into the
chip-loading deck, follow the steps below (Fig. 6.7).
(a) Align a metal frame on a micropillar chip by holding the metal frame at an
angle of 45° against the edge of the micropillar chip (Fig. 6.7A). The dents
at the bottom of the metal frame should be perfectly aligned with the circles
at the bottom of the micropillar chip.
(b) Slowly drop the metal frame so that all the micropillars go into the holes on
the metal frame (Fig. 6.7B).
(c) Gently insert the micropillar chip with the metal frame into the mag-
netic micropillar chip zig by holding the side of the chip with two fin-
gers (Fig. 6.7B).
(d) The micropillar chip will be laid flat and held firmly while scanning due to
strong magnets in the micropillar chip zig (Fig. 6.7C). Note: The strong
magnetic attraction between the zig and the metal frame may impose
severe force on the micropillar chip. Therefore, make sure that all the
micropillars go into the holes on the metal frame properly before placing
it on the zig.
10. Open the sliding door and load the micropillar zig with the chips into the chip-
loading deck by inverting the zig (Fig. 6.8). Note: The symbol (4) on the zig
faces the same direction as the symbol on the chip-loading deck.
11. In case of scanning microwell chips, the chip should be sealed with a gas-
permeable membrane to prevent water evaporation in the microwells and spill-
ing of the sample over the objective lens.
12. A microwell chip zig consists of two parts—part I is an open metal frame for
placing microwell chips and part II is a magnetic frame that can hold the
microwell chips firmly while scanning due to strong magnets (Fig. 6.9).
(A) (B)
(C) (D)
Fig. 6.9 (a) A microwell chip inserted in the part I microwell chip zig, (b) The front side of the
part II microwell chip zig, (c) The part I zig combined with the part II zig, and (d) The back side
of the part II zig combined with the part I zig. The symbol (4) on the back of the part II zig high-
lighted in red should be aligned with the bottom of the microwell chips inserted.
13. In order to place microwell chips on the microwell chip zig and load it into the
chip-loading deck, follow the steps below (Fig. 6.9).
(a) Place and insert microwell chips in the part I microwell chip zig (Fig. 6.9A)
so that the back side of the microwell chip is exposed for scanning while
the front side of the chip is sealed with the gas-permeable membrane, fac-
ing the part II microwell chip zig (Fig. 6.9B).
(b) Take the part II zig and place it on the part I zig so that the symbol [4] on
the part II zig is aligned with the bottom of the microwell chips (Fig. 6.9C, D).
Note: The microwell chip will be laid flat and held firmly while scanning
due to strong magnets in the microwell chip zig.
14. Click ‘Open’ in the ‘Chip layout’ tool bar (Fig. 6.10). A window displaying a
list of chip files will appear (Fig. 6.11). Note: These chip files contain XYZ
coordinates for the micropillar chip and the microwell chip at different objec-
tive lenses.
6.3 Protocols 133
15. Select a desired chip file from the list by clicking ‘Open’ in the ‘Chip layout’
window (Fig. 6.11). Note: Refer to parameter setting procedures to set and
change XYZ coordinates in a given chip file.
16. In the ‘Home’ window, click the ‘Spot position’ icon in the ‘XYZ Move’ tab to
enable accurate selection of individual spots (i.e., micropillars or microwells)
for live view (Fig. 6.12).
17. Select the number of chips being scanned from the ‘Chip selection’ tab in the
‘Home’ window (Fig. 6.12).
18. In the ‘Live view’ tab, click the ‘Live’ icon to observe a real-time cell image and
check an optimum exposure time by dragging the cursor left to right in the
‘Exposure time’ section or manually entering the value in the ‘Exposure time’
section (Fig. 6.12).
134 6 3D-Cultured Cell Image Acquisition
Fig. 6.13 ‘Histogram’ window displaying the histogram of the red-fluorescent image from a red
filter. The red intensity is higher than 255, meaning that the exposure time is a bit too high
19. Click the ‘Histogram’ icon in the ‘Home’ window to enable the display of fluo-
rescence histogram (Fig. 6.12).
20. To find a proper exposure time, use the ‘Histogram’ tool bar to analyze the
wavelength and also visually inspect the real-time cell image. An exposure time
which gives a histogram between 80 and 90 % of maximum pixel range
(i.e., 255) for the given channel is considered as an optimum exposure
time (Fig. 6.13). Note: Do not select too high exposure time as photobleach-
ing of fluorescence might occur at higher exposure time.
21. In the ‘Region of interest’ tab, click ‘ROI’ icon to display the region of interest
(ROI) in the image window and use the ‘Radius’ and ‘Margin’ cursor or enter
the value to increase and decrease the size of the ROI (Fig. 6.12).
22. In the ‘Chip scan’ tool bar, click ‘Set’ to activate the setting window for each
chip. Assign a label for each chip in the ‘Folder’ section, set the exposure time
in ‘Exposure’ section, and select the desired filter from the drop-down list in the
‘Filter’ section. (Fig. 6.14). Note: A single chip can be scanned using four dif-
ferent filters and exposure setting depending on the user requirement. Make
sure that the rest of the drop-down list in ‘Filter’ section is set to ‘UNUSED’
when scanning the chip with less than four filters setting. Select ‘Multiband’
filter for acquiring images from cells stained with multiple fluorescent dyes.
The drop-down list in ‘Auto-Exp’ field is usually set to ‘UNUSED’ when
scanning the chip with a user-defined exposure setting.
6.3 Protocols 135
23. Select the magnification of an objective lens used for image acquisition from the
drop down list in the ‘Lens’ field and select the number of steps for autofocus
from the drop down list in the ‘Steps’ field in the ‘Chip scan’ tab (Fig. 6.12). Note:
Select a larger number of steps when the chip is bent or not placed flat. It will
cover a larger distance in the Z direction to obtain an optimum image in focus.
24. Click the ‘Scan’ icon to begin chip scanning.
25. The images obtained from individual filters are saved in multiple folders with
the name of the specific filter and the exposure time used for that filter in the
format of ‘filter_exposure time’ (e.g., Multiband_150). These folders are stored
inside another folder with the name of the chip provided by users (Refer to Step
22). The user-named folders for multiple chips are stored in a big folder which
is labeled in the format of ‘year-month-day’ (e.g., 2016-06-01).
26. After the scanning is done, remove the chips from the scanner and store the
scanned chips in the dark until disposed.
27. Close the scanner software and turn off the computer, the S+ Scanner, and the
light source (Fig. 6.3). Note: To switch off the light source, first press the ‘ON/
OFF’ button in the front until the blue LED turns off and the countdown
begins from 300 on the digital display. Wait until the countdown ends and
then turn off the main switch of the light source located on the back.
6.3.2.1 etting the Position of Filters and the Distance of Each Step
S
for Autofocus
This tool bar displays all the list of filters available for ‘Live’ view of cells and chip
scanning, allows to set and change the position of the filters, and also allows to
modify the distance of each step for autofocus (Refer to Step 22). Note: The num-
ber of steps and the distance of each step are important to obtain an optimum
image in focus using the autofocus function. The S+ Scanner takes multiple
136 6 3D-Cultured Cell Image Acquisition
pictures in the Z direction but save only one best-looking image. Ideally, the dis-
tance of each step should be equal to the height of cell spheroids.
1. Click ‘Move’ button to select the desired filter, which has predetermined excita-
tion and emission spectra for image acquisition (Fig. 6.15 and Appendix). For
example, clicking the ‘Move’ button for the ‘Orange’ filter enables ‘Live’ view
of cells/spheroids stained with orange/red fluorescent dyes in the excitation/
emission range of the ‘Orange’ filter.
2. Filters are set to the position displayed on the ‘Position’ column by default.
Change or reset the position of individual filters only when the filter position is
misplaced due to machine malfunctioning or user errors.
3. To change or reset the filter position, click the ‘Set’ button to activate the ‘Filter
setting’ window and enter the position of a respective filter so that ‘Live’ cell
image is visible uniformly without distortion (Fig. 6.15). Note: The filter posi-
tion has to be increased or decreased by few units only if needed. For example,
the ‘Multiband’ filter position has to be increased or decreased to 23 or 21,
respectively.
4. The ‘Excitation’ and ‘Emission’ columns are used to calculate and set the dis-
tance of each step for autofocus (Fig. 6.15). Note: It doesn’t indicate actual
excitation and emission wavelengths. When we select the number of ‘Steps’ in
the ‘Chip scan’ tab for autofocus (Fig. 6.12), it uses the distance of each step
(i.e., the thickness of one step) calculated from the following formula.
ë û
where the numerical apertures for the 4× lens and the 20× lens are 0.13 and 0.5,
respectively.
Note: The recommended distance of each step is 10–20 μm (e.g., 400 nm emission
wavelength with 4× lens) for 3D-cultured cells and 2–5 μm (e.g., 120 nm emission
wavelength with 4× lens) for 2D cell monolayers. For a small step distance, it will
require a large number of ‘Steps’ for better autofocus. The emission wavelength
setting can be changed for each filter used and the morphology of cells.
The ‘Chip layout’ tool bar allows to design the scanning layout (i.e., the number of
micropillars or microwells per chip) and determine specific XYZ coordinates of left
top (LT), right top (RT), and left bottom (LB) corners of the chip for accurate image
acquisition.
1. Click ‘Open’ in the ‘Chip layout’ tool bar to select a desired chip file from the
list (Fig. 6.16).
2. Select a desired chip file and click ‘Open’ in the window. For example, the layout
of the micropillar chip at 4× magnification lens is selected (Fig. 6.17).
3. To change the X, Y, and Z coordinates of the chips in the desired chip file, click
‘Set/Save’ button among four chips loaded (i.e., Chip 1, Chip 2, Chip 3, and
Chip 4) (Fig. 6.16). Note: This will allow one to change the X, Y, and Z coordi-
nates of LT, RT, and LB corners of the chip selected.
4. To read accurate XYZ coordinates, click the ‘Axis’ icon in the ‘XYZ Move’ tab
in ‘Home’ window to open the ‘Axis’ window (Fig. 6.18).
5. In the ‘Live view’ tab, click the ‘Live’ icon to observe a real-time cell image.
Note: We need to place three micropillars or microwells at the LT, RT, and LB
corners in the center of the ROI.
Fig. 6.18 ‘Home’ window showing the ‘Axis’ icon highlighted in red in the ‘XYZ Move’ tab
6.4 Summary
Imaging technology is the key determinant for a successful HCI assay. An ideal HCI
system enables the user to acquire high-resolution images of fluorescently labeled
cells at a high speed in real time. In this chapter, we introduced some commonly
available HCI devices and provided simple procedures for acquiring high-content
images in a high-throughput fashion from the micropillar/microwell chip platform
using the S+ Scanner. These procedures can be further modified to acquire high-
content images from traditional HTS platforms such as 96- and 384-well plates.
6.5 Appendix
http://www.semrock.com/SetDetails.aspx?id=2777
140 6 3D-Cultured Cell Image Acquisition
2. Orange filter
http://www.semrock.com/SetDetails.aspx?id=2799
3. Blue filter
http://www.semrock.com/SetDetails.aspx?id=2779
References 141
4. Green filter
http://www.omegafilters.com/Products/Curvomatic
References
1. Buchser, W., Collins, M., Garyantes, T., Guha, R., Haney, S., Lemmon, V., et al. (2012). Assay
development guidelines for image-based high content screening, high content analysis and
high content imaging. In S. GS, C. NP, H. Nelson, M. Arkin, D. Auld, C. Austin, et al. (Eds.),
Assay guidance manual. (pp. 1–69). Bethesda, MD: Eli Lilly & Company and the National
Center for Advancing Translational Sciences.
2. Talbot, C. B., McGinty, J., Grant, D. M., McGhee, E. J., Owen, D. M., Zhang, W., et al. (2008).
High speed unsupervised fluorescence lifetime imaging confocal multiwell plate reader for
high content analysis. Journal of Biophotonics, 1(6), 514–521. doi:10.1002/jbio.200810054.
3. Celli, J. P., Rizvi, I., Blanden, A. R., Massodi, I., Gidden, M. D., Pogue, B. W., et al. (2014).
An imaging-based platform for high-content, quantitative evaluation of therapeutic response in
3D tumour models. Scientific Reports. doi:10.1038/srep03751.
4. Romano, S. N., & Gorelick, D. A. (2014). Semi-automated imaging of tissue-specific fluores-
cence in Zebrafish Embryos. Journal of Visualized Experiments, 87, e51533. doi:10.3791/51533.
5. Bickle, M. (2010). The beautiful cell: High-content screening in drug discovery. Analytical and
Bioanalytical Chemistry, 398, 219–226. doi:10.1007/s00216-010-3788-3.
142 6 3D-Cultured Cell Image Acquisition
6. Jahr, W., Schmid, B., Schmied, C., Fahrbach, F. O., & Huisken, J. (2015). Hyperspectral light
sheet microscopy. Nature Communications, 6, 1–7. doi:10.1038/ncomms8990.
7. Reynaud, E. G., Peychl, J., Huisken, J., & Tomancak, P. (2015). Guide to light-sheet micros-
copy for adventurous biologists. Nature Methods, 12(1), 30–34. doi:10.1038/nmeth.3222.
8. Rae Chi, K. (2012). High on high content: A guide to some new and improved high-content
screening systems. The Scientist Magazine, 1–7.
9. Starkuviene, V., & Pepperkok, R. (2007). The potential of high-content high-throughput
microscopy in drug discovery. British Journal of Pharmacology, 152(1), 62–71. d oi:10.1038/
sj.bjp.0707346.
Chapter 7
High-Content Image Analysis
Sean Yu, Pranav Joshi, Dong Woo Lee, and Moo-Yeal Lee
Contents
7.1 Introduction...................................................................................................................... 143
7.2 Materials.......................................................................................................................... 145
7.2.1 ImageJ.................................................................................................................. 145
7.2.2 S+ Chip Analysis................................................................................................. 145
7.3 Protocols.......................................................................................................................... 146
7.3.1 3D Cell Image Analysis with ImageJ.................................................................. 146
7.3.2 Examples of Image Processing with ImageJ....................................................... 149
7.3.2.1 Hue Filter............................................................................................ 149
7.3.2.2 Background Subtraction, Brightness Filter, and Region
of Interest (ROI................................................................................... 150
7.3.2.3 Outlier Exclusion................................................................................ 151
7.3.2.4 The Performance of the Plugin........................................................... 151
7.3.3 Image Deconvolution........................................................................................... 152
7.3.3.1 Performance of Deconvolution........................................................... 154
7.3.4 Plotting Dose Response Curves with S+ Chip Analysis...................................... 154
7.4 Summary.......................................................................................................................... 159
References................................................................................................................................. 159
7.1 Introduction
lent overview [4]. In this chapter, we summarize protocols for analyzing images of
3D-cultured cells on the micropillar/microwell chip obtained by staining with
multiple fluorescent dyes.
7.2 Materials
ImageJ is used to extract fluorescent intensity from 3D-cultured cell images, and S+
Chip Analysis is used to plot dose response curves from the extracted fluorescence
data in high throughput.
7.2.1 ImageJ
7.3 Protocols
Fig. 7.1. Selection of the BioPrinting plugin from the Plugins menu
7.3 Protocols 147
the other dye to get separate quantitative data for two separate dyes in the same
image set. As for the hue range, calcein AM, for example, has a working range
of 20–130 (Table 7.1).
148 7 High-Content Image Analysis
Table 7.1. Hue ranges of fluorescent dyes with green, red, and blue colors.
Dyes Fluorescent color Minimum hue Maximum hue
Calcein AM Green 20 130
Ethidium homodimer-1 Red 0 50
Hoechst 33342 Blue 120 200
Tetramethyl rhodamine methylester Red 0 50
(TMRM)
14
Block Fluorescence
Block 1 1
Block 2 2
Block 3 3
532
38 Block 4 4
Block 5 5
Block 6 6
(A) (B)
Fig. 7.4. (a) The *.scn data output file that matches the layout of the micropillar/microwell chip
(14 × 38 data array) and (b) The *.txt output is in 6 block format (1 × 532 data array) for ease of
use elsewhere such as statistical software.
Note: The following images were selected for their undesirable features to demon-
strate the performance of the BioPrinting plugin and are not representative of images
acquired from the micropillar/microwell chip platform using the S+ Chip Scanner.
Fig. 7.5. The performance of hue filtering: (a) Original image with cells (in green), dust (in red)
and artifacts (in blue) and (b) Hue filtered image with cells alone. Green dots represent live cells
on the micropillar chip
150 7 High-Content Image Analysis
When the hue of the stained cells is similar to background fluorescence, dust, and artifacts,
we can use several methods to eliminate unwanted noise fluorescence. Brightness filter
(Fig. 7.3) simply reduces pixel brightness values below a certain threshold to zero
(Fig. 7.6). This method is particularly useful for images with significant base fluores-
cence. Background subtraction (Figs. 7.3 and 7.7) can be used as an alternative of, or
in conjunction with, the brightness filter in reducing background fluorescence via the
rolling ball algorithm [5]. Often times, the undesirable fluorescence signals can be
Fig. 7.6. The effect of brightness filter: (a) Original image with no background filter applied, (b)
Brightness filter 10 applied, (c) Brightness filter 30 applied, and (d) Brightness filter 50 applied.
Blue dots represent nucleus stained with Hoechst 33342
Fig. 7.7. The effect of background subtraction: (a) Original image with no rolling background
subtraction applied and (b) Rolling background subtraction applied
7.3 Protocols 151
Fig. 7.8. Typical image processing: (a) Original image with edge artifacts, (b) Brightness filtered
image, (c) Background subtracted image, and (d) Region of interest (ROI) for fluorescence extrac-
tion. Blue dots represent nucleus stained with Hoechst 33342
found outside the area of the micropillar/microwell where cells are seeded and grown.
In that case, we can specify a region of interest (ROI) and have the software only
extract fluorescence data from those regions (Fig. 7.8).
Without any processing, the scanned images often have high background fluores-
cence and high variability. By applying the various aforementioned features of the
plugin, we can significantly reduce the unwanted background fluorescence and
152 7 High-Content Image Analysis
diminish the inflated variability as shown in the following images. Briefly, the
micropillar chips were seeded with varying cell concentrations and stained with 2
different dyes—fluo-4 acetoxymethyl ester (AM), and monochlorobimane
(mBCl)—to determine the relationship between fluorescence and cell seeding den-
sity. Unprocessed raw fluorescence data acquired from the scanned micropillar
chips were compared with the processed data (Fig. 7.10). Both Fluo-4 AM and
mBCl had high background fluorescence, which was significantly reduced by the
plugin. Variability caused by spot detachment, dust, or artifacts was also signifi-
cantly reduced—for example, see mBCl at 2 million cells/mL.
Image processing is a well-developed field, and there are vast and increasing num-
ber of solutions that could be borrowed and implemented to enhance signals while
eliminating noises. One such method is deconvolution. When cells are encapsulated
in a 3D gel, some of them may not be in focus at the same time. As a result, the out-
of-focus cell images appear somewhat poor and blurred, which could impact the
resulting fluorescence data. Mathematically, blurring can be considered as a convo-
lution between the in-focus (theoretical) image and a blurring factor. The blurring
factor could be removed using deconvolution so that an un-blurred image can be
remained. Therefore, we are focusing on image deconvolution as a potential solu-
tion to the blurring problem. Various image processing toolboxes or suites exist that
can perform deconvolution such as the image processing toolbox for MATLAB or
the several deconvolution plugins for ImageJ. We decided to use Image Restoration
software developed by Advanced Technology Inc. (ATI) for its decoupling and cus-
tomizability of kernel estimation methods and deconvolution methods. The Kernel
estimation method refers to the process of identifying the blurring factor, and the
deconvolution algorithm uses the estimated kernel to deconvolve the images. The
blind form of Richardson and Lucy (RL) algorithm combined with Inter-Level
Intra-Level Deconvolution (ILILD) was determined to be ideal for images of cells
7.3 Protocols 153
3.5 1.4
Fluorescence (Million)
2.5 1.0
2.0 0.8
1.5 0.6
1.0 0.4
0.5 0.2
0.0 0.0
0 2 4 6 8 10 0 2 4 6 8 10
(A) Seeding density (Million cells/mL) (B) Seeding density (Million cells/mL)
10 3.0
2.5
Fluorescence (Million)
Fluorescence (Million)
8
2.0
6
1.5
4
1.0
2
0.5
0 0.0
0 2 4 6 8 10 0 2 4 6 8 10
(C) Seeding density (Million cells/mL) (D) Seeding density (Million cells/mL)
Fig. 7.10. The effect of the BioPrinting plugin (hue filter, background subtraction, brightness fil-
ter, ROI, and outlier exclusion) on cell images from the microarray chip: (a) Unfiltered data of cells
stained with fluo-4 AM, (b) Filtered fluo-4 AM data, (c) Unfiltered data from cells stained with
mBCl, and (d) Filtered mBCl data
Fig. 7.11. ATI’s Image Restoration software for deconvolution. Blue dots represent nucleus
stained with Hoechst 33342
3. Select the ‘Deconvolution’ method from the drop-down list. Note: ILILD is
recommended.
4. Click on the ‘Batch test’ button and select the directory where the images are stored.
5. Click ‘Execute’ to run the software for deconvolution.
3D cell cultures, especially multi-layered 3D cell cultures in the microwell chip, can
be challenging to image due to the cells not being in the same focal plane due to
relatively large spotting volume (typically 350–1000 nL). To test the efficacy of
deconvolution on multi-layered 3D cell cultures, microwell chips were seeded with
a double layer of hoechst 33342-stained cells in varying concentrations, imaged,
and processed using the ImageJ Bioprinting plugin as well as the deconvolution
software. Deconvolution was found to increase the sensitivity and the slope of the
fluorescence calibration as well as the wellness of fit (Fig. 7.12).
1. Open S+ Chip Analysis program by double clicking the S+ Chip Analysis icon
on the desktop (Fig. 7.13).
2. Click the ‘Experiment’ button on the ‘Menu’ bar to open the ‘Experiment
condition’ window (Fig. 7.14).
7.3 Protocols 155
Original
5
Fluorescence (Million)
RL - ILILD
4
Fig. 7.12. The effect of deconvolution on fluorescence calibration curves. The blind Richardson
and Lucy (RL) kernel estimation method and the Inter-Level Intra-Level Deconvolution (ILILD)
algorithm were applied to the images of double-layered Hoechst 33342-stained cells
Fig. 7.14. S+ Chip Analysis window with the ‘Menu’ bar and the ‘Experiment condition’
window
3. Click the ‘Browse’ button in the ‘Experiment condition’ window and select the
experiment folder where image data are saved (Fig. 7.15).
4. Click the ‘Show’ button to see the data. The data will be displayed on a popup
window (Fig. 7.15).
5. Select the ‘Chip Platform’ for the layout of spots (i.e., experimental conditions)
on the chip from the drop down list (Fig. 7.16).
6. Enter the dilution factor in the ‘Compound Dilution’ field (Fig. 7.16).
7. Enter the name of the cells used in the ‘Cell Type’ field, select the unit of com-
pound measurement (e.g., μM) from drop down list in the ‘Unit’ field, and then
click the ‘Apply’ button (Fig. 7.16).
156 7 High-Content Image Analysis
Fig. 7.15. The ‘Experiment condition’ window with browse button for selecting the desired folder
and the ‘Show’ button for displaying the data, both highlighted in red
Fig. 7.17. The ‘Status’ window displaying the status of data analysis
8. The cell type and the unit of test compounds will be applied to all the data and
will appear in the ‘Cell Type’ and ‘Unit’ grid.
9. Enter the compound names and the highest concentrations tested against the
cell type in the first and second row, respectively (Fig. 7.16). Note: The name
of six compounds and the respective concentrations should be entered
individually in column numbered from 1 to 6, respectively.
10. Enter the name of enzymes or fluorescent dyes or any reagents applied to the
chip in the ‘Enzyme’ grid (Fig. 7.16). Note: While analyzing multiple chips, if
the compound names, concentration values, and enzymes/fluorescent dye
names are identical, the user can copy and paste parameters from one chip to
other chips.
11. If outliers are to be detected and removed automatically, check the ‘Detect
Outliers’ option (Fig. 7.16).
12. After entering the parameters press the ‘Save’ button to save the parameters
(Fig. 7.16).
7.3 Protocols 157
13. Press ‘Run’ button to start analyzing the data (Fig. 7.16). Status of analysis will
be displayed in ‘Status’ window (Fig. 7.17).
14. After the data is analyzed, ‘Graph plotting completed’ message will appear.
Press OK to complete the analysis (Fig. 7.18).
15. The ‘Experiment Result’ window will show up after the completion of data
analysis. Inspect the individual dose response curve and check ‘Select All’ to
select all the graphs obtained or select individual graphs from the chip prior to
combining the replicates from different chips (Fig. 7.19).
16. Select ‘Export’ in the menu bar to export the result/graph into an MS-Word file
(Fig. 7.19). Note: This process allows one to combine dose response curves
from multiple replicate chips and draw one dose response curve.
17. Once the exporting process is complete, a ‘Save As’ window will appear
with ‘Report.doc’ as its default file name. Press save if you want to save it
with the default file name or else enter a desired file name and press ‘Save’
button (Fig. 7.20).
18. A ‘Select Plot Types’ window will show up, check ‘% Viability vs. Log concen-
tration—summary plot’ and click ‘Export’ button (Fig. 7.21).
19. A pop up window with a message ‘Document file created’ will appear. Press
OK to finish exporting the file (Fig. 7.22).
20. A summary table with test conditions and IC50 values as well as corresponding
dose response curves will be saved in the same folder containing data file
(Fig. 7.23). Note: To produce a conventional sigmoidal dose-response curve,
the fluorescence intensities of all cell spots are normalized to the fluorescence
intensity of 100 % live cell spots (e.g., cell spots contacted with no compound)
158 7 High-Content Image Analysis
Fig. 7.20. The ‘Save As’ window appears to save the result in a desired filename inside a desired
folder
Fig. 7.23. Examples of dose-response curves with IC50 values obtained from data analysis in S+
Chip Analysis software. Dose-response curves of lovastatin for (a) nucleus morphology, (b) mito-
chondrial membrane potential, and (c) glutathione level are presented
and then plotted against the logarithm of test compound concentrations. The
sigmoidal dose-response curves (variable slope) and IC50 values for each test
condition are obtained using the following equation:
(
Y = Bottom + é( Top - Bottom ) / 1 + 10(
ë
LogIC50 - X )* H
)ùû
where IC50 is the midpoint of the curve, H is the hill slope, X is the logarithm
of test concentration, and Y is the response (% live cells), starting at Bottom and
going to Top with a sigmoid shape.
7.4 Summary
In this chapter, we introduced protocols for processing the images of 3D-cultured cells
on the micropillar/microwell chip in a high-throughput manner using ImageJ. We also
demonstrated a deconvolution algorithm for deblurring the out of focus images using
Image Restoration software developed by Advanced Technology Inc. (ATI), South
Korea. Finally, we explained step-by-step protocols for plotting dose-response curves
and calculating IC50 values using S+ Chip Analysis software, specifically designed for
visualizing and analyzing images from microarray chip platforms.
References
1. Lang, P., Yeow, K., Nichols, A., & Scheer, A. (2006). Cellular imaging in drug discovery.
Nature Reviews. Drug Discovery, 5(4), 343–356. doi:10.1053/j.gastro.2006.06.028.
2. Jahr, W., Schmid, B., Schmied, C., Fahrbach, F. O., & Huisken, J. (2015). Hyperspectral light
sheet microscopy. Nature Communications, 6, 1–7. doi:10.1038/ncomms8990.
3. Scherf, N., & Huisken, J. (2015). The smart and gentle microscope. Nature Biotechnology,
33(8), 815–818. doi:10.1038/nbt.3310.
160 7 High-Content Image Analysis
4. Eliceiri, K. W., Berthold, M. R., Goldberg, I. G., Ibáñez, L., Manjunath, B. S., Martone, M. E.,
et al. (2012). Biological imaging software tools. Nature Methods, 9(7), 697–710. doi:10.1038/
nmeth.2084.
5. Sternberg, S. (1983). Biomedical image processing. IEEE Computer, 16(1), 22–34.
6. Fish, D. A., Brinicombe, A. M., Pike, E. R., & Walker, J. G. (1995). Blind deconvolution by
means of the Richardson–Lucy algorithm. Journal of the Optical Society of America. A, 12(1),
58–65.
7. Y. Ding, I. Park, X. Cui, Van Huan. Nguyen, Hakil Kim, Trung Dung Do et al. (2015). Inter-
level and intra-level deconvolution based image deblurring algorithm for wide field micros-
copy. In 2015 8th International Conference on Biomedical Engineering and Informatics
(BMEI). Shenyang: IEEE (pp. 90–95). doi:10.1109/BMEI.2015.7401479.
Chapter 8
Applications of Microarray Bioprinting
Contents
8.1 Introduction...................................................................................................................... 161
8.2 Assay Development for Microarray Bioprinting Technologies....................................... 162
8.2.1 Hepatotoxicity Assays......................................................................................... 162
8.2.1.1 Phase I and Phase II Drug Metabolizing Enzyme Assays.................. 164
8.2.1.2 Drug Transporter Assays.................................................................... 165
8.2.1.3 Oxidative Stress Assays...................................................................... 166
8.2.2 Neurotoxicity Assays........................................................................................... 166
8.2.2.1 Oxidative Stress and Related Assays.................................................. 167
8.2.2.2 Ion Channel Assays............................................................................ 167
8.2.2.3 Drug Metabolism Assays.................................................................... 168
8.3 Simulation of the In Vivo Microenvironment: Liver Applications.................................. 169
8.4 Other Applications........................................................................................................... 171
8.5 Summary.......................................................................................................................... 171
References................................................................................................................................. 172
8.1 Introduction
Adverse drug reactions (ADRs) are among the five leading causes of death in the
United States [1]. Among ADRs, hepatotoxicity and cardiotoxicity are the most
common events observed. Other organ-specific ADRs, including neurotoxicity and
renal toxicity, are observed in the presence of certain drugs. Additionally, the cost
for successful drug development can range from $160 million to $2.5 billion, and it
takes ten to 15 years from lead compound discovery to clinical evaluation [2]. About
60 % of this cost comes from clinical trials [3]. With roughly a post-marketing fail-
ure rate of one drug per year, the human, financial, and capital tolls are further
increased.
One of the major issues with drug development is the reliance on animal models
to predict ADRs before clinical trials are conducted. Animal models have been
shown to be poor predictors for ADRs, as other mammals do not adequately mimic
all the human tissues and organ functions [4–7]. Additionally, animal models are
generally significantly more expensive than preliminary in vitro toxicity studies.
Thus, it is a goal of scientists to develop a suite of in vitro toxicity assays that can
be used to accurately predict human in vivo toxicity, and potentially reduce the costs
associated with drug development.
In this section, we will focus on assays that can be implemented for in vitro tox-
icity studies on the microarray printing platform. Specific focus will be placed on
hepatotoxicity assays, focusing on phase I and phase II drug metabolizing enzymes
(DMEs), transporter assays, and oxidative stress detection. Additionally, neurotox-
icity will be explored in regards to assays that detect specific mechanisms for neu-
rotoxicity, and how this toxicity is differentiated between adult neuronal cells and
neural stem cells (NSCs).
The liver is arguably the most important organ tied to toxicity assays as the liver is
responsible for clearance of toxic substances from the body, and hepatotoxicity is
the most common form of ADR that is associated with market withdrawals. Toxicity
in the liver can stem from general elevated exposures to certain compounds, known
as intrinsic drug-induced liver injury (DILI), or it can stem from individual varia-
tions in the tissue microenvironment and the genetic of the individual, known as
idiosyncratic drug-induced liver injury (IDILI). Here, we will focus on the assays
associated with detection of DILI.
Drug metabolism in the liver is illustrated in Fig. 8.1. An external, apical membrane
transporter can allow drugs to enter the cell, though it is also possible for the
structure of the molecule to pass through the membrane via diffusion. These com-
pounds are then covalently modified in 2–3 steps to be transported into the bile via
8.2 Assay Development for Microarray Bioprinting Technologies 163
Drug
Passive
diffusion
Influx
transporter
(NTCP, OCT,
OATP) Necrosis/
Protein Apoptosis Nucleus
Lipid
Parent peroxidation damage
drug ER stress
DNA
Phase I metabolism damage
(CYP, FMO, MAO) ROS/RNS
generation NFkB, Nrf2 Cell
Reactive Antioxidants survival
metabolite (GSH, SOD,
catalase, GPx)
Phase II metabolism Protein
(UGT, SULT, GST, NAT) adduct
ROS
MPT,
Conjugated Necrosis
mtDNA
metabolite (RIPK, PGAM) ATP/glucose damage
depletion
Apoptosis ROS
Efflux (JNK, caspase, Bcl2) Mitochondria
transporter
(BSEP, BCRP,
MDR, MRP)
Fig. 8.1. Simplified mechanisms of metabolism-mediated drug toxicity in liver cells. Abbreviations
are used as follows: NTCP Na+-taurocholate cotransporting polypeptid, OCT organic cation trans-
porter, OATP organic anion transporting polypeptide, BSEP bile salt export pump, BCRP breast
cancer resistance protein, MDR multidrug resistance protein, MRP multidrug resistance-associated
protein, CYP cytochrome P450, FMO flavin-containing monooxygenase, MAO monoamine oxi-
dase, UGT UDP-glucuronosyltransferases, SULT sulfotransferase, GST glutathione S-transferase,
NAT N-acetyl transferase, ROS reactive oxygen species, RNS reactive nitrogen species, NFκB
nuclear factor kappa-light-chain-enhancer of activated B cells, Nrf2 nuclear factor erythroid
2-related factor 2, GSH glutathione, SOD superoxide dismutase, GPx glutathione peroxidase, MPT
mitochondrial pore transition, mtDNA mitochondrial DNA, RIPK receptor-interacting serine/
threonine protein kinase, PGAM phosphoglycerate mutase, JNK c-Jun N-terminal kinase, Bcl2
B-cell lymphoma 2, APC antigen-presenting cell, KC Kupffer cell
the effects of damaging DNA, proteins, and lipids. Mechanisms that clear ROS
and RNS from the body can contribute to limiting the effects of ADRs, but if the
activities of the enzymes that combat oxidative stress are compromised, this can
also have a deleterious effect on hepatocytes. Finally, molecules damaged by any
forms of the drug or ROS/RNS can be presented on the surface of the hepatocyte,
potentially triggering an acute inflammatory response. While most responses are
immediate and short term, sometimes the inflammatory response is sustained and
this can lead to further damage, or potentially autoimmunity against the liver, which
is potentially fatal.
properly assess CYP450 activity. Additionally, there are substrates that exist which
can be metabolized by isolated CYP450s, but may not be adequate to permeate the
cell membrane. Human liver microsomes (HLMs) may be used to convey or alter
CYP450 activity to cells lacking CYP450 activity, or can be used as a suitable con-
trol for assaying activity of certain compounds with various CYP450s [9]. HLM
activity can be assayed using microarray bioprinting technology, which offers the
alternative of conventional 96-well plate assays at a lower cost.
Phase II drug metabolism is the replacement of reactive functional groups added
by phase I drug metabolism with bulkier, easily cleared functional groups. Phase II
enzymes include uridine 5′-diphosphoglucuronosyltransferases (UGTs), sulfotrans-
ferases (SULTs), glutathione S-transferases (GSTs), and N-acetyltransferases
(NATs). For all of these enzymes, there exist clinically relevant polymorphisms that
make certain individuals more susceptible to ADRs. Specific protein assays that are
relevant to ADRs have been developed for measuring UGT and GST activity. UGT
activity can be measured by using 4-methylumbelliferone (4-MU), a fluorescent sub-
strate of the coumarin family that undergoes glucuronidation in the presence of UGT
[12, 13]. Additionally, owing to the role of UGT in clearance of steroids, β-estradiol
can be used to assay UGT1A1 activity via detection of estradiol 3-glucuronide [14].
GST can be measured and distinguished using 1-chloro-2,4-dinitrobenzene (CDNB)
as a substrate [15]. This reaction involves the removal of the chlorine via reaction
with the sulfur on glutathione (GSH). While CDNB can be used as a substrate for
total GST activity, it is measured via absorbance, and less ideal for microarray bio-
printing applications. Monochlorobimane (MCB) is recommended as the reaction
with GSH produces a fluorescent compound (excitation 380 nm/emission 461 nm),
which has been used previously to assess GSH levels in cells.
While oxidative stress can affect numerous organelles, antioxidant enzymes local-
ized to the mitochondria are the antioxidants most associated with individualized
ADRs. This makes the mitochondrion the primary indicator for liver functionality
tests related to oxidative stress. For general mitochondrial toxicity assays, tetra-
methyl rhodamine (TMRM) can be used to stain for measuring mitochondrial mem-
brane potential, as depolarization of the membrane is a symptom of apoptosis [23].
Cytochrome c release and oxygen consumption rate (OCR) may also be used to
determine effects on energetic metabolism [24]. OCR determination utilizes the
metabolic turnover of O2 in energetic metabolism [25, 26]. OCR can be measured
fluorescently using a variety of methods, though fluorescent probes are most com-
mon. MitoXpress from Luxcel (ex/em 380/650) can be used for direct fluorescence
measurement of OCR by quantifying oxidative phosphorylation, though time-
resolved fluorescence (TR-F) is recommended instead of direct fluorescence.
Similar assays are available for OCR measurements from other companies.
If antioxidant levels are innately low, GSH may also be used to determine
effects of drug on mitochondria. MCB may be used as a fluorogenic substrate to
quantify GSH levels [27], but the activity of GSH as a function of antioxidant
stress must be distinguished from that of GSH used in phase II drug metabolism.
While GSH acts as a general indicator for oxidative stress, this can be sufficient
enough to indicate dysfunction with antioxidant enzymes, or even the general
presence of an ADR.
Aside from hepatotoxicity, ADRs are also found in neural cell lineages, particularly
neural stem cells (NSCs) [28]. NSCs are the cells that eventually differentiate into
the various lineages associated with the central nervous system, including neurons,
oligodendrocytes, and astrocytes. Much of the toxicity that drives ADRs in NSCs
relates to damaging or blocking ion channels on the cell membrane surface, induc-
ing oxidative stress, denaturing of nucleic acids and proteins from drug metabolites,
8.2 Assay Development for Microarray Bioprinting Technologies 167
Another way ADRs are triggered in NSCs involve the blocking of ion channels. Ion
channels are associated with maintaining plasma membrane potential, but they can
also be found on the membranes of various organelles. Within axonal differentiated
168 8 Applications of Microarray Bioprinting
Fig. 8.2. The conversion of CoroNa green via intracellular esterases to a fluorescent indicator for
intracellular Na+ ions
neural cell lineages and NSCs, ion channels also play a role in synaptic transmission
and the generation of action potentials, which is a common behavior associated with
cells of the CNS and all versions of muscle cells. Disruption in ion channels may
lead to disruption in the transmittance of action potentials, and the subsequent loss
of signals. K+, Na+, Ca2+, and Cl− may all be blocked within neural cells and NSCs.
The disruption in ion flow could result from competitive ionic interactions of ions
with similar valence charges (i.e., transition elements Ni2+ and Cd2+ with Ca2+),
covalent modification of proteins, or competitive binding with necessary ligands
that may allow channels to open and close. Common compounds used as ion chan-
nel blocker controls include 4-aminopyridine, disopyramide, amlodipidine besyl-
ate, and 4,4′-diisothiocyanostilbene-2,2′-disulfonic acid (DIDS).
There are commercially available assays used to detect ion channel activity and ion
channel blocking. FluxOR potassium ion channel assay from Life Technologies uses
thallium ions to trigger the opening of K+ channels, activating the fluorescent dye inside
the cell, while also using probenecid to prevent the exit of the fluorescent dye. The
CoroNa green sodium indicator from Life Technologies uses a fluorescein based mol-
ecule with a binding pocket for Na+, which is activated inside cells despite Na+ ion
concentrations generally being greater extracellularly (Fig. 8.2). Fluo-4AM reagent
from Life Technologies is believed to detect Ca2+ molecules intracellularly via cleavage
of the acetoxymethyl ester groups. This may be coupled with probenecid to maintain
the fluorescent moieties inside the cell. Premo halide sensor from Life Technologies
uses I− to mimic the influx of Cl− ions. I− will quench a yellow fluorescent agent inside
the cell. It is known so far that Fluo-4AM is compatible with a high-throughput bio-
printing system, and that the rest of these reagents are cell compatible.
Another such mechanism for detecting NSC toxicity is through drug metabolism
itself. Like hepatocytes, adult neural cells and NSCs metabolize drugs and can be
adversely affected by mutations in DMEs [34]. Drug metabolism in NSCs utilizes a
8.3 Simulation of the In Vivo Microenvironment: Liver Applications 169
similar process of phase I and phase II metabolism, and there are separate classes of
transporters for compound influx and efflux. CYP450s are the DMEs primarily
responsible for phase I drug metabolism in NSCs, and there is some overlap with
the individual CYP450 isoforms that are responsible for drug metabolism.
Additionally, phase II DME of the GST family have been tied to individual ADRs,
and dysfunction in transporter MDR1 may also increase susceptibility to ADRs.
NSC assays which deal with drug metabolism focus on similar assays that are
involved in assessing DME-related toxicity in hepatocytes. While the protocols are
identical for detection of activity with these enzymes, many of the CYP450 iso-
forms unique to NSCs are involved in the metabolism of fatty acids and steroids [35,
36]. Thus, substrates for metabolism and assaying activity of neuronal CYP450
enzymes can be such biological compounds, or their analogues. Likewise, many of
the GST and MDR1 assays used on hepatocytes are also compatible with NSCs,
including CDNB and MCB for GST activity, and verapamil for MDR1. These
assays are compatible with cells and are fluorescent, making them suitable assays to
be used for determining toxicity on a microarray chip platform.
Automatic liquid dispensation for microarray technologies has until recently seen
limited use outside the organism. While research has focused on development of
whole 3D tissue constructs, technology has not reached the point where whole
organs can be printed for transplantation purposes [37]. The creation of physio-
logical networks is quite complex, and maintaining and mimicking all of the nec-
essary functions of a 3D-bioprinted tissue is difficult. However, microarray 3D
bioprinting can be used to mimic small scale in vivo behaviors. While this does
not make it suitable for large scale cultures, miniaturization of the bioprinting
process can still give insight into certain in vivo behaviors in a high-throughput
manner. In this section, we will focus on the use of bioprinting 3D cell cultures for
high-throughput applications, with a focus on co-culture systems related to mimicking
the liver (Fig. 8.3).
In order to understand the nature of co-culture systems, we must also understand
the cells and vasculature associated with the tissue that is to be mimicked. In the
case of livers, there are four types of cells that make up over 95 % of the cell volume
and count: hepatocytes, Kupffer cells (KCs), sinusoidal endothelial cells, and
hepatic stellate cells [38, 39]. Hepatocytes make up the majority of the cell volume
and count (roughly 60 % of the cell count and 80 % of the cell mass) [38]. These
cells serve a myriad of functions, including drug metabolism, bile flow, cholesterol
synthesis, glycolysis, gluconeogenesis, and amino acid catabolism [39]. The roles
the hepatocytes take on relate closely towards their proximity towards the portal
vein, the main supplier of oxygen to the liver [39]. The other cell types are localized
to other parts of the liver. The KCs are the resident macrophages of the liver that
initiate inflammation in response to toxic injury [38, 40, 41]. Sinusoidal endothelial
170 8 Applications of Microarray Bioprinting
Bioprinting with S+
microarray spotter
Combinatorial cell printing Chip
into the microwell chip In vitro In vivo scanning
study study
Chip incubation
in growth media Chip
implanting
Cell staining
Cells mixed with Chip
“bioink” Photocrosslinking recovery
in a 96-well plate
Image
analysis
Fig. 8.3. Microarray three-dimensional (3D) bioprinting technology for creating human tissues
by dispensing multiple human cells in biomimetic hydrogels (‘bioinks’) layer-by-layer with preci-
sion printing robots, thereby potentially revolutionizing tissue engineering and disease modeling
for screening therapeutic drugs and studying toxicology. Miniaturized liver tissue blocks (as small
as 1 mm3) are generated by printing several layers of human hepatic cells in photocrosslinkable
hydrogels with extracellular matrices (ECMs) and growth factors (GFs) into a microwell chip
using S+ MicroArrayer. After gelation, the microwell chip containing hundreds of biomimetic
conditions is incubated in a petri dish with growth media or implanted in immunocompromised
animals for rapidly testing optimum microenvironments to create biomimetic human tissues.
Created human liver tissue constructs are stained and scanned with S+ Scanner for high-content
imagining (HCI) of hepatic functions as well as predictive assessment of drug toxicity
cells border the blood vessels, and are important as they have interactions with any
blood cells (including KCs) and can present antigens to CD4+ (mature) helper T
cells [42, 43]. They are also important on limiting the effects of shear on hepato-
cytes [38, 39, 44]. Finally, hepatic stellate cells exist in the space of disse, a region
between the endothelial cells, and the hepatocytes. The stellate cells store fatty acids
and some vitamins, but they may behave like fibroblasts and synthesize collagen if
liver restructuring is necessary [38, 39, 45].
To mimic the behavior of the liver, focus has shifted towards 3D co-culture sys-
tems, utilizing both hepatocytes and non-parenchymal cells. While KCs are most
frequently incorporated into these systems owing to their role in response towards
ADRs, the other cell types have been used too [46, 47]. Model systems that have
been used to create this environment include large scale hepatic bioreactor [48, 49],
microfluidic technologies [50–52], layered-sheets of fibroblasts to initiate growth
conditions [53], and 3D hydrogel technologies for microarrays [54]. For the drug
discovery process, including testing for ADRs, 3D hydrogel microarrays are ideal
because of their high throughput nature. Hepatic bioreactors cannot be scaled down
at low cost, microfluidics can provide some replicates, but not with the same effi-
ciency as a microarray can, and fibroblast systems could potentially be more com-
plicated than the hydrogels, and many really do not adequately mimic the 3D liver
microenvironment. Therefore, it is more desirable to use a microarray system for
drug development testing.
8.5 Summary 171
8.5 Summary
References
1. Reuben, A., Koch, D. G., & Lee, W. M. (2010). Drug-induced acute liver failure: Results of a
U.S. multicenter, prospective study. Hepatology, 52(6), 2065–2076. doi:10.1002/hep.23937.
2. Morgan, S., Grootendorst, P., Lexchin, J., Cunningham, C., & Greyson, D. (2011). The cost of
drug development: A systematic review. Health Policy (Amsterdam, Netherlands), 100(1),
4–17. doi:10.1016/j.healthpol.2010.12.002.
3. Preventable adverse drug reactions: A focus on drug interactions. 2016. Retrieved from http://
www.fda.gov/Drugs/DevelopmentApprovalProces.
4. Elliott, N. T., & Yuan, F. A. N. (2011). A review of three-dimensional in vitro tissue models for
drug discovery and transport Studies. Journal of Pharmaceutical Sciences, 100(1), 59–74.
doi:10.1002/jps.22257.
5. Godoy, P., Hewitt, N. J., Albrecht, U., Andersen, M. E., Ansari, N., Bhattacharya, S., et al.
(2013). Recent advances in 2D and 3D in vitro systems using primary hepatocytes, alternative
hepatocyte sources and non-parenchymal liver cells and their use in investigating mechanisms
of hepatotoxicity, cell signaling and ADME. Archives of Toxicology, 87(8), 1315–1530.
doi:10.1007/s00204-013-1078-5.
6. Kimlin, L. C., Casagrande, G., & Virador, V. M. (2013). In vitro three-dimensional (3D)
models in cancer research: An update. Molecular Carcinogenesis, 52(3), 167–182. doi:10.1002/
mc.21844.
7. Keogh, J. P. (2012). Membrane transporters in drug development. Advances in Pharmacology,
63, 1–42. doi:10.1016/B978-0-12-398339-8.00001-X.
8. Lin, J., Schyschka, L., Mühl-Benninghaus, R., Neumann, J., Hao, L., Nussler, N., et al. (2012).
Comparative analysis of phase I and II enzyme activities in 5 hepatic cell lines identifies Huh-7
and HCC-T cells with the highest potential to study drug metabolism. Archives of Toxicology,
86(1), 87–95. doi:10.1007/s00204-011-0733-y.
9. Walsky, R. L., & Obach, R. S. (2004). Validated assays for human cytochrome P450 activities.
Drug Metabolism and Disposition, 32(6), 647–660. doi:10.1124/dmd.32.6.647.
10. Zarowna-Dabrowska, A., McKenna, E. O., Schutte, M. E., Glidle, A., Chen, L., Cuestas-Ayllon,
C., et al. (2012). Generation of primary hepatocyte microarrays by piezoelectric printing.
Colloids and Surfaces B: Biointerfaces, 89(1), 126–132. doi:10.1016/j.colsurfb.2011.09.016.
11. Meli, L., Jordan, E. T., Clark, D. S., Linhardt, R. J., & Dordick, J. S. (2012). Influence of a
three-dimensional, microarray environment on human cell culture in drug screening systems.
Biomaterials, 33(35), 9087–9096. doi:10.1016/j.biomaterials.2012.08.065.
12. Leite, S. B., Iwona, W.-Z., Zaldivar, J. M., Airola, E., Reis-Fernandes, M. A., Mennecozzi, M.,
et al. (2012). 3D HepaRG model as an attractive tool for toxicity testing. Toxicological
Sciences, 130(1), 106–116.
13. Leite, S. B., Teixeira, A. P., Miranda, J. P., Tostões, R. M., Clemente, J. J., Sousa, M. F., et al.
(2011). Merging bioreactor technology with 3D hepatocyte-fibroblast culturing approaches:
Improved in vitro models for toxicological applications. Toxicology in Vitro, 25(4), 825–832.
doi:10.1016/j.tiv.2011.02.002.
References 173
14. Chang, J. H., Plise, E., Cheong, J., Ho, Q., & Lin, M. (2013). Evaluating the in vitro inhibition
of UGT1A1, OATP1B1, OATP1B3, MRP2, and BSEP in predicting drug-induced hyperbiliru-
binemia. Molecular Pharmaceutics, 10(8), 3067–3075.
15. Liang, Q., Sheng, Y., Jiang, P., Ji, L., Xia, Y., Min, Y., et al. (2011). The gender-dependent
difference of liver GSH antioxidant system in mice and its influence on isoline-induced liver
injury. Toxicology, 280(1–2), 61–69. doi:10.1016/j.tox.2010.11.010.
16. Kis, E., Ioja, E., Rajnai, Z., Jani, M., Méhn, D., Herédi-Szabó, K., et al. (2012). BSEP inhibi-
tion: In vitro screens to assess cholestatic potential of drugs. Toxicology In Vitro, 26(8), 1294–
1299. doi:10.1016/j.tiv.2011.11.002.
17. Lee, J. K., Marion, T. L., Abe, K., Lim, C., Pollock, G. M., & Brouwer, K. L. R. (2010).
Hepatobiliary disposition of troglitazone and metabolites in rat and human sandwich-cultured
hepatocytes: Use of Monte Carlo simulations to assess the impact of changes in biliary excre-
tion on troglitazone sulfate accumulation. The Journal of Pharmacology and Experimental
Therapeutics, 332(1), 26–34. doi:10.1124/jpet.109.156653.tion.
18. Cascorbi, I., & Haenisch, S. (2010). Pharmacogenetics of ATP-binding cassette transporters
and clinical implications. Methods in Molecular Biology, 596, 95–121.
doi:10.1007/978-1-60761-416-6.
19. Fromm, M. F. (2004). Importance of P-glycoprotein at blood-tissue barriers. Trends in
Pharmacological Sciences, 25(8), 423–429. doi:10.1016/j.tips.2004.06.002.
20. Stieger, B. (2011). The role of the Sodium-Taurocholate Cotransporting Polypeptide (NTCP)
and of the Bile Salt Export Pump (BSEP) in physiology and pathophysiology of bile forma-
tion. In M. F. Fromm & R. B. Kim (Eds.), Drug transporters. Berlin: Springer.
doi:10.1007/978-3-642-14541-4.
21. Goral, V. N., Hsieh, Y.-C., Petzold, O. N., Clark, J. S., Yuen, P. K., & Faris, R. A. (2010).
Perfusion-based microfluidic device for three-dimensional dynamic primary human hepato-
cyte cell culture in the absence of biological or synthetic matrices or coagulants. Lab on a
Chip, 10(24), 3380–3386. doi:10.1039/c0lc00135j.
22. Thompson, R. A., Isin, E. M., Y, L., Weidolf, L., Page, K., Wilson, I., et al. (2012). In vitro
approach to assess the potential for risk of idiosyncratic adverse reactions caused by candidate
drugs. Chemical Research in Toxicology, 25(8), 1616–1632. doi:10.1021/tx300091x.
23. Xu, J. J., Diaz, D., & O’Brien, P. J. (2004). Applications of cytotoxicity assays and pre-lethal
mechanistic assays for assessment of human hepatotoxicity potential. Chemico-Biological
Interactions, 150(1), 115–128. doi:10.1016/j.cbi.2004.09.011.
24. Porceddu, M., Buron, N., Roussel, C., Labbe, G., Fromenty, B., & Borgne-Sanchez, A. (2012).
Prediction of liver injury induced by chemicals in human with a multiparametric assay on
isolated mouse liver mitochondria. Toxicologic Pathology, 129(2), 332–345.
25. Hynes, J., Swiss, R. L., & Will, Y. (2012). High-throughput analysis of mitochondrial oxygen
consumption. In C. M. Palmeira & A. J. Moreno (Eds.), Mitochondrial bioenergetics: Methods
and protocols (Vol. 810, pp. 103–117). New York: Humana Press.
doi:10.1007/978-1-61779-382-0.
26. Aleo, M. D., Luo, Y., Swiss, R., Bonin, P. D., Potter, D. M., & Will, Y. (2014). Human drug-
induced liver injury severity is highly associated with dual inhibition of liver mitochondrial
function and bile salt export pump. Hepatology, 60(3), 1015–1022. doi:10.1002/hep.27206.
27. Ong, M. M. K., Latchoumycandane, C., & Boelsterli, U. A. (2007). Troglitazone-induced
hepatic necrosis in an animal model of silent genetic mitochondrial abnormalities. Toxicological
Sciences, 97(1), 205–213. doi:10.1093/toxsci/kfl180.
28. Fernandes, T. G., Kwon, S. J., Bale, S. S., Lee, M. Y., Diogo, M. M., Clark, D. S., et al. (2010).
Three-dimensional cell culture microarray for high-throughput studies of stem cell fate.
Biotechnology and Bioengineering, 106(1), 106–118. doi:10.1002/bit.22661.
29. Fernandes, T. G., Kwon, S. J., Lee, M. Y., Clark, D. S., Cabral, J. M. S., & Dordick, J. S.
(2008). On-chip, cell-based microarray immunofluorescence assay for high-throughput analy-
sis of target proteins. Analytical Chemistry, 80(17), 6633–6639. doi:10.1021/ac800848j.
30. Kwon, S. J., Lee, D. W., Shah, D. A., Ku, B., Jeon, S. Y., Solanki, K., et al. (2014). High-
throughput and combinatorial gene expression on a chip for metabolism-induced toxicology
screening. Nature Communications, 5, 3739. doi:10.1038/ncomms4739.
174 8 Applications of Microarray Bioprinting
31. Ulasov, I., Nandi, S., Dey, M., Sonabend, A. M., & Lesniak, M. S. (2011). Inhibition of Sonic
hedgehog and Notch pathways enhances sensitivity of CD133+ glioma stem cells to temozolo-
mide therapy. Molecular Medicine, 17(1–2), 103–112. doi:10.2119/molmed.2010.00062.
32. Tegenge, M. A., Rockel, T. D., Fritsche, E., & Bicker, G. (2011). Nitric oxide stimulates
human neural progenitor cell migration via cGMP-mediated signal transduction. Cellular and
Molecular Life Sciences, 68(12), 2089–2099. doi:10.1007/s00018-010-0554-9.
33. Blurton-jones, M., Spencer, B., Michael, S., Castello, N. A., Agazaryan, A. A., Davis, J. L.,
et al. (2014). Neural stem cells genetically-modified to express neprilysin reduce pathology in
Alzheimer transgenic models Neural stem cells genetically-modified to express neprilysin
reduce pathology in Alzheimer transgenic models. Stem Cell Research & Therapy, 5(2), 46.
34. Farrel, K., Joshi, P., Roth, A., Kothapalli, C. R., & Lee, M.-Y. (2016). High-throughput screen-
ing of toxic chemicals against neural stem cells. In J. L. Sherley (Ed.), Human stem cell toxi-
cology (pp. 31–63). Cambridge: The Royal Society of Chemistry.
35. Ferguson, C. S., & Tyndale, R. F. (2011). Cytochrome P450 enzymes in the brain: Emerging
evidence of biological significance. Trends in Pharmacological Sciences, 32(12), 708–714.
doi:10.1016/j.tips.2011.08.005.
36. Miksys, S., & Tyndale, R. F. (2013). Cytochrome P450-mediated drug metabolism in the
brain. Journal of Psychiatry & Neuroscience, 38(3), 152–163. doi:10.1503/jpn.120133.
37. Murphy, S. V., & Atala, A. (2014). 3D bioprinting of tissues and organs. Nature Biotechnology,
32(8), 773–785. doi:10.1038/nbt.2958.
38. Arias, I., Wolkoff, A., Boyer, J., Shafritz, D., Fausto, N., Alter, H., & Cohen, D. (2011).
The liver: Biology and pathobiology. Chichester: Wiley.
39. Boron, W. F., & Boulpaep, E. L. (2009). Medical physiology (2nd ed.). Philadelphia, PA:
Saunders Elsevier.
40. Jaeschke, H. (2011). Reactive oxygen and mechanisms of inflammatory liver injury: Present
concepts. Journal of Gastroenterology and Hepatology, 26(Suppl 1), 173–179.
doi:10.1111/j.1440-1746.2010.06592.x.
41. Zimmermann, H. W., Trautwein, C., & Tacke, F. (2012). Functional role of monocytes and
macrophages for the inflammatory response in acute liver injury. Frontiers in Physiology, 3,
1–18. doi:10.3389/fphys.2012.00056.
42. Kim, Y., & Rajagopalan, P. (2010). 3D hepatic cultures simultaneously maintain primary hepa-
tocyte and liver sinusoidal endothelial cell phenotypes. PloS One, 5(11), 1–10. doi:10.1371/
journal.pone.0015456.
43. Sato, Y., Tsukada, K., & Hatakeyama, K. (1999). Role of shear stress and immune responses
in liver regeneration after a partial hepatectomy. Surgery Today, 29(1), 1–9.
44. Tuleuova, N., Lee, J. Y., Lee, J., Ramanculov, E., Zern, M. A., & Revzin, A. (2010). Using
growth factor arrays and micropatterned co-cultures to induce hepatic differentiation of
embryonic stem cells. Biomaterials, 31(35), 9221–9231. doi:10.1016/j.
biomaterials.2010.08.050.
45. Ueno, T., Sata, M., Sakata, R., Torimura, T., Sakamoto, M., Sugawara, H., et al. (1997).
Hepatic stellate cells and intralobular innervation in human liver cirrhosis. Human Pathology,
28(8), 953–959.
46. Messner, S., Agarkova, I., Moritz, W., & Kelm, J. M. (2013). Multi-cell type human liver
microtissues for hepatotoxicity testing. Archives of Toxicology, 87(1), 209–213. doi:10.1007/
s00204-012-0968-2.
47. Kostadinova, R., Boess, F., Applegate, D., Suter, L., Weiser, T., Singer, T., et al. (2013). A
long-term three dimensional liver co-culture system for improved prediction of clinically rel-
evant drug-induced hepatotoxicity. Toxicology and Applied Pharmacology, 268(1), 1–16.
doi:10.1016/j.taap.2013.01.012.
48. Fiegel, H. C., Kneser, U., Kluth, D., & Rolle, U. (2010). Hepatic tissue engineering.
Handchirurgie, Mikrochirurgie, Plast Chir Organ der Deutschsprachigen Arbeitsgemeinschaft
für Handchirurgie Organ der Deutschsprachigen Arbeitsgemeinschaft für Mikrochirurgie der
Peripher Nerven und Gefässe Organ der Vereinigung der Deut, 42(6), 337–41. doi:10.105
5/s-0030-1252045.
References 175
49. Hoekstra, R., Nibourg, G. A., van der Hoeven, T. V., Plomer, G., Seppen, J., Ackermans, M. T.,
et al. (2013). Phase 1 and phase 2 drug metabolism and bile acid production of HepaRG cells
in a bioartificial liver in absence of dimethyl sulfoxide. Drug Metabolism and Disposition,
41(3), 562–567. doi:10.1124/dmd.112.049098.
50. Bhatia, S. N., & Ingber, D. E. (2014). Microfluidic organs-on-chips. Nature Biotechnology,
32(8), 760–772. doi:10.1038/nbt.2989.
51. Toh, Y.-C., Lim, T. C., Tai, D., Xiao, G., van Noort, D., & Yu, H. (2009). A microfluidic 3D
hepatocyte chip for drug toxicity testing. Lab on a Chip, 9(14), 2026–2035. doi:10.1039/
b900912d.
52. Bale, S. S., Vernetti, L., Senutovitch, N., Jindal, R., Hegde, M., Gough, A., et al. (2014). In
vitro platforms for evaluating liver toxicity. Experimental Biology and Medicine (Maywood,
N.J.), 239(9), 1180–1191. doi:10.1177/1535370214531872.
53. Nakazawa, K., & Shinmura, Y. (2011). Effects of culture conditions on a micropatterned co-
culture of rat hepatocytes with 3T3 cells. Journal of Bioprocessing & Biotechniques, 01(3).
doi:10.4172/2155-9821.S3-002.
54. Li, C. Y., Stevens, K. R., Schwartz, R. E., Alejandro, B. S., Huang, J. H., & Bhatia, S. N.
(2014). Micropatterned cell-cell interactions enable functional encapsulation of primary hepa-
tocytes in hydrogel microtissues. Tissue Engineering. Part A, 20(617), 2200–2212. doi:10.1089/
ten.TEA.2013.0667.
55. Ma, Y., Ji, Y., Huang, G., Ling, K., Zhang, X., & Bioprinting, X. F. (2015). 3D cell-laden
hydrogel microarray for screening human periodontal ligament stem cell response to extracel-
lular matrix. Biofabrication, 7(4), 044105. doi:10.1088/1758-5090/7/4/044105.
56. Bailey, S. N., Sabatini, D. M., & Stockwell, B. R. (2004). Microarrays of small molecules
embedded in biodegradable polymers for use in mammalian cell-based screens. Proceedings
of the National Academy of Sciences of the United States of America, 101(46), 16144–16149.