You are on page 1of 31

REVIEW

High-throughput screening assays for the


identification of chemical probes
James Inglese, Ronald L Johnson, Anton Simeonov, Menghang Xia, Wei Zheng, Christopher P Austin & Douglas S Auld

High-throughput screening (HTS) assays enable the testing of large numbers of chemical substances for activity in diverse areas
of biology. The biological responses measured in HTS assays span isolated biochemical systems containing purified receptors or
enzymes to signal transduction pathways and complex networks functioning in cellular environments. This Review addresses
factors that need to be considered when implementing assays for HTS and is aimed particularly at investigators new to this field.
We discuss assay design strategies, the major detection technologies and examples of HTS assays for common target classes,
cellular pathways and simple cellular phenotypes. We conclude with special considerations for configuring sensitive, robust,
informative and economically feasible HTS assays.

A variety of strategies (Fig. 1) have been used for HTS assays, includ- information can guide active compound selection by differentiating
ing the measurement of catalytic activity from a purified enzyme1, selectivity among related targets13, or by distinguishing the inhibition
a reconstituted complex2,3, a cellular extract4,5 or a phenotype6–9 in intact of a cellular pathway from a cytotoxic response in a cell-based assay14–16.
cells. Configuring assays to function within the constraints imposed by Second, is the stimulus dependent only on the compound being tested,
HTS differentiates an HTS assay from traditional laboratory assays, as or is the response to the compound conditioned by another stimulus,
outlined in Table 1. The 96-well microtiter plate, originally designed to such as a fully efficacious concentration of agonist (when screening for
facilitate serological studies in virology10, was later recognized as a suit- an antagonist), a subefficacious concentration of agonist (when screen-
able replacement for cuvette, tube and dish-based assays. Combined with ing for a potentiator) or a permissive temperature? In conditional assays,
spectrophotometric plate readers, the microtiter plate quickly became compound activities independent of the conditional stimulus can be
the standard for performing parallel assays for rapid preliminary eval- false positives or can act outside the pathway of interest (Fig. 1d). Third,
uation of compound bioactivity. During the 1990s, the tremendous does the duration of the response occur within seconds upon application
increase in chemical libraries and assays with sensitivities amenable to of the stimulus, thereby requiring a rapid read (for example, a calcium
miniaturization further drove the parallel processing concept that today transient), or within minutes or hours of the stimulus, thereby permit-
is practiced in the 96-, 384- or 1,536-microwell plate formats, among ting endpoint or multiple time point measurements? These parameters
others. ‘High throughput’ is a relative term, but it is generally defined as will define the basic format of the primary HTS.
the testing of 10,000 to 100,000 compounds per day, accomplished with Follow-up assays, also referred to as ‘secondary screens’ or ‘counter
mechanization that ranges from manually operated workstations to fully screens’, should be planned before the HTS. It is essential to view the
automated robotic systems11,12. Assays designed for the purpose of HTS primary HTS as the initial step of an integrated process. Follow-up tests
attempt to integrate biological fidelity with enabling assay and screening validate compounds targeting the intended biological interaction and
technologies. Subsequently, understanding the pharmacological impact assist in the elimination of compounds that generate a positive signal
of the assay design is critical to the identification of biologically relevant via other mechanisms. False positives can be anticipated based on the
compounds from HTS. assay design (for example, fluorescent compounds in an assay with
a fluorescent readout) or eliminated by testing in an independent or
Assay design for HTS ‘orthogonal’ assay that uses a different methodology or biological read-
An effective HTS strategy considers both the primary and subsequent out of target activity. As an example, cyclic AMP (cAMP), an important
secondary assay designs carefully. First, is the nature of the response to cellular second messenger, can be measured with an output response
be measured clearly defined: does the signal increase, decrease, change signal either directly or indirectly proportional to cellular [cAMP]
in nature or location or belong to part of a more complex response? Can (Fig. 2). cAMP assays showing inversely proportional responses to
the response be measured via a single parameter, or is a multiparametric changes in [cAMP] in principle can function in an HTS-confirmatory
output possible? For example parameters that provide counter-screen assay sequence in which selected candidates identified in the HTS are
then tested for initial validation of activity. Such ‘preliminary confirma-
US National Institutes of Health Chemical Genomics Center, National Institutes tion’ strategies using HTS assays in sequence (or parallel) can greatly
of Health, 9800 Medical Center Drive, Bethesda, Maryland 20892-3370, USA. lower the retesting of false positives in substantially lower throughput
Correspondence should be addressed to J.I. (jinglese@mail.nih.gov). secondary assays that are generally more complex and designed to be a
Published online 18 July 2007; doi:10.1038/nchembio.2007.17 more physiological measure of the biological system under study17.

466 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY


REVIEW

performance18. The Z factor’s unitless scale


PEP ADP
(ranging from 0 to 1) allows comparison of
different assays and screens using the control
wells (Z') and sample wells (Z) of the plate.
b Comparison of Z' and Z in assay validation has
Pyruvate ATP + luciferin been used to determine compound screening
a concentration, as assays in which Z' > Z pre-
dict percent of actives greater than the <1%

target value for single concentration screening
GT
P-γ
S paradigms. If compounds are absent from the
sample field (for example, DMSO only), a low
GPCR Z score may be diagnostic of a faulty reagent

dispense or contamination problem in the
sample wells, for example.
An equally significant component of assay
or validation is assay sensitivity—the accuracy
hν P GFP and reproducibility of potency measurements
of control compounds or conditions. The
GTP-γS minimum significance ratio (MSR, Table 2) is
GFP hν
Eu used to measure the reproducibility of potency
P
values and defines the statistically significant
potency range that can be measured in an
c assay19,20. During HTS, the MSR can be calcu-
d lated from titrations of control compounds on
some or all assay plates in the screen to track
assay sensitivity variation, which is often an
indicator of reagent stability.

Detection of biological responses in HTS


The photon, which is created by either a fluo-
rescent or luminescent mechanism, is the
Figure 1 Assay categories and methodologies. (a) An isolated enzyme catalyzes turnover of a principal output in HTS assays. As the most
profluorescent blue- or red-emitting substrate. (b) A coupled-enzyme assay in which the ATP product, widely used HTS detection modality (Table 3;
generated by the pyruvate kinase–catalyzed reaction (ribbon structure), is detected by a “reporter” see Supplementary Table 1 online for addi-
enzyme (firefly luciferase) to create a luminescent output (hν). (c) A membrane preparation is used to tional detail), fluorescent energy is highly
measure the activity of a GPCR through the ligand-dependent activation of associated G proteins as tunable, which allows the development of a
measured by the binding of a fluorescent europium (Eu3+)-labeled GTP-γS analog. (d) A phenotypic
variety of fluorescent sensors21. Fluorescence
assay designed to probe a signaling pathway dependent on coactivation of an extracellular receptor
(blue) and an intracellular protein (yellow). The intracellular protein is activated by a pharmacological can be bright and occur on a timescale ranging
agent (, blue dashed arrow) as a prerequisite to revealing potentiators (compounds that are active from 10−9 s (for typical organic fluorophores)
only in the presence of the costimulus) of the pathway, as measured by the nuclear translocation of to 10−4 s (for lanthanide cryptates), thereby
a GFP-labeled protein. Such potentiators could act by inhibiting proteins that repress the pathway allowing for light’s many optical properties
(black dashed arrows). to be exploited by a number of detection
methods22. For example, the application of
plane-polarized light in fluorescence polar-
Regardless of the approach taken, the HTS assay protocol must be ization (FP) can be used to measure a probe’s rotational perturba-
simplified from its original form (Table 1), and the following param- tions, thereby enabling simple assay designs dependent on a single
eters must be optimized: (i) sufficient sensitivity to enable the identifi- labeled ligand (ref. 23, and ref. 24 and other articles in the same issue).
cation of compounds with low potency or efficacy. (ii) Reproducibility Alternatively, time-resolvable fluorescent trivalent lanthanides such as
and stability of the biological response between wells and plates. This Eu3+ and Tb3+ can be used to increase assay sensitivity by suppress-
parameter is dependent on both assay reagents and hardware (such as ing background fluorescence from assay components and library
dispensers, incubators and detectors). (iii) Accuracy of the positive and compounds25,26. Further, fluorescent probes can transfer energy to
negative controls as compared with the known pharmacology of the tar- other chromophores or fluorophores in assays that measure distance-
get. (iv) Economic feasibility (frequently measured as cost per well). dependent fluorescence or fluorescence quenching via resonance energy
Validation of an assay protocol determines its suitability for HTS and transfer (FRET or QFRET) mechanisms25–27.
involves two parts: performance and sensitivity18 (Fig. 3). Several sta- For cellular assays in particular, chemical28,29 and genetically encoded27
tistical parameters are used to evaluate assay performance. Signal win- fluorescent probes allow customized, multiplexed or direct measurement
dow and Z factor (Table 2) are calculations that gauge the fold response of events from biological components present at low concentrations30.
between maximum and minimum output signals and the precision of Coupled with scanning microplate cytometers31 or microscope-based
this response within a plate and across plates. The signal window mea- imaging systems, these probes are moving HTS from population-aver-
sures the statistically significant difference between the maximum and aged outputs toward the enumeration of individual cellular events8,9,
minimum signal, but it is not as reliable as Z factor for predicting assay thereby enhancing the information content obtainable from each well.

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 467


REVIEW

Table 1 Differences in allowed parameters between laboratory “bench top” and HTS assays
Parameter Bench top HTS
Protocol May be complex with numerous steps, aspirations, washes Few (5–10) steps, simple operations, addition only preferred
Assay volume 0.1 ml to 1 ml <1 µla to 100 µl
Reagents Quantity often limited, batch variation acceptable, may be unstable Sufficient quantity, single batch, must be stable over prolonged period
Reagent handling Manual Robotic
Variables Many—for example, time, substrate/ligand concentration, compound, Compoundb, compound concentration
cell type
Assay container Varied—tube, slide, microtiter plate, Petri dish, cuvette, animal Microtiter plate
Time of measurement Milliseconds to months Minutes to hours
Measurements as endpoint, multiple time points, or continuous Measurements typically endpoint, but also pre-read and kinetic
Output formats Plate reader, radioactivity, size separation, object enumeration, Plate reader—mostly fluorescence, luminescence and absorbance
images interpreted by human visual inspection
Reporting format “Representative” data; statistical analysis of manually curated dataset Automated analysis of all data using statistical criteria
aSpecial reagent dispensers required. bIdeally available in milligram quantity with analytical verification of structure and purity.

The absence of excitation light energy in the generation of a lumi- owing to the short path length that results from low volumes, though
nescence signal greatly reduces compound interference in HTS (Fig. 4). several assays have been implemented in 1,536-well plates42,43.
Although the signal strength can be significantly lower than fluores- Variations on the use of these technologies in the design of assays are
cence, the background is negligible, which allows luminescence-based impressive, and a discussion of each is beyond the scope of this paper.
assays to have an enormous dynamic range, on the order of 1,000 times Assays suitable for HTS comprise a subset of these and must meet the
more sensitive than equivalent fluorescence-based assays32. Both che- requirements outlined in Table 1. The next section highlights the well-
miluminescence and bioluminescence use a chemical reaction to create exploited assay formats.
light. In the case of bioluminescence, an enzyme (typically from the
firefly Photinus pyrali) generates light through its action on a luciferin Assays for specific targets
substrate, analogs of which have been used to develop assays for cyto- Historically, screening assays have been performed on a limited set of
chrome P450s, proteases and monoamine oxidases32. Further, the active targets, many involving purified proteins. Approximately two thirds of
site of the P. pyrali luciferase has been mutated to spectrally vary the therapeutic targets are comprised of enzymes and receptors44. More
luminescence33, thereby allowing the use of multiple luminescent sen- recently, HTS labs are increasing attention on cell-based high-content
sors in an assay15. screens (HCS) that analyze biological events at subcellular resolution8,9; a
Luciferase reporters are prevalent in HTS assays because of convenient 2006 survey indicated that roughly half of HTS assays are cell-based and
detection, ample sensitivity and their ubiquitous occurrence in academic that the use of HCS is rapidly growing, particularly in secondary screen-
research. Dual luciferase reporter assays with distinct substrate speci- ing45. The types of proteins for which drugs have been successfully devel-
ficities, kinetics or emission maxima have been useful for identifying oped are relatively restricted. About half of experimental and marketed
activities specific to the signaling pathway of interest. Frequently, the drugs target five main protein families: G protein–coupled receptors
second reporter is used to normalize gene expression or discriminate (GPCRs), kinases, proteases, nuclear receptors (NRs) and ion channels46.
cytotoxic compounds34,35. Luciferase reporters can be combined with Because of the great interest in these protein classes for drug develop-
other detection formats as well, for instance in tandem with a green ment, many approaches and technologies (Table 3 and Supplementary
fluorescent protein (GFP) reporter14, β-galactosidase or Alamar blue16 Table 1) have been developed to assay these targets, and in some cases
to assess cytotoxicity. they are broadly applicable to other target classes as well.
The β-lactamase reporter offers particular advantages for HTS. For
example, the β-lactam–coupled coumarin-fluorescein substrate, CCF4/ Enzymes. Enzymes are comprised of six categories: oxidoreductases,
AM, allows a ratiometric read that helps minimize variation in cell num- transferases, hydrolases, lyases, isomerases and ligases. They catalyze a large
ber or substrate concentration because the emission maxima of the cleaved number of chemical transformations. In cases in which the enzyme, cofac-
and intact β-lactamase substrates (460 and 530 nm, respectively) are dis- tor and substrate can be obtained in an isolated, active form, the devel-
tinct (reviewed in ref. 36). In addition, the uncleaved substrate can be opment of an HTS assay generally hinges on the availability of reagents
used to assess cytotoxicity, as fluorescence is detected only if the substrate and technologies to enable substrate or product detection. Here, protein
is taken up by living cells and de-esterified37,38. kinases and proteases will serve to illustrate various approaches to HTS
A hybrid system involving both bioluminescence and fluorescence assay design for enzymes.
is represented by BRET-based assays39 (Table 3 and Supplementary
Table 1). Here luciferase luminescence stimulates a proximal fluorescent Protein kinases. Protein kinases mediate the most prevalent post-transla-
protein emission, thereby bypassing the need for exogenous excitation tional modification in cells, phosphorylate an array of targets in signaling
light. However, this technique may be more challenging to optimize for cascades and cycles, and are important for therapeutic intervention, par-
HTS assays because of lower S:B ratios and higher background from ticularly oncology47. In this section we discuss biochemical approaches to
weakly associated target proteins40. protein kinase assays, as cell-based assays are covered in the later sections
In addition, radioactive decay in conjunction with the scintillation on cellular signaling. The biochemical approaches can be divided into
proximity assay (SPA) format is applicable to a host of targets41, and two categories: generic assays independent of subfamily (for example,
low-cost absorbance assays have proven reliable in measuring analytes tyrosine versus serine/threonine) and antibody-based formats that detect
(for example, H+ or Pi), cellular staining or cell growth. However, in an epitope within the phosphorylated product. The latter are generally
microtiter plates, absorbance measurements can have low S:B ratios used for a particular subfamily or specific protein kinase.

468 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY


REVIEW

Protein kinases catalyze phosphoryl group transfers from ATP FP. Such assays are advantageous in targeting kinase activity dependent
to hydroxyl groups of amino acid side chains of serine, threonine or on, for example, hierarchical phosphorylation or kinase cascades55. Also,
tyrosine. An easily implemented assay format that has been used suc- multiplexed readouts of kinase activity and cellular phosphoproteins are
cessfully in HTS detects ATP depletion via luciferase, an ATP-depen- possible with these formats56.
dent photoenzyme48,49. A disadvantage of this approach is that 50% A unique system that assays the kinase inactive state relies on ATP
conversion of the ATP is required to achieve a two-fold S:B. This may binding and uses enzyme fragment complementation (EFC) for detec-
require increased enzyme levels for sufficient signal, thus decreasing tion57. Here the general kinase inhibitor, staurosporine, is linked to a
sensitivity and increasing reagent costs. Alternatively, kinase activity enzyme donor (ED) fragment of β-galactosidase that when displaced
can be measured through ADP production. One assay format uses an from the kinase ATP pocket restores β-galactosidase activity through α
ADP-specific antibody and labeled ADP to generate an FP competition complementation with the acceptor enzyme (EA) component58.
binding assay to monitor ADP product formation50. A coupled enzyme A final general point, illustrated with kinases, relates to assay designs
strategy for ADP detection that yields a fluorescent readout has also that recapitulate as closely as possible the native target to uncover new
been described51. mechanisms of compound action. As most biochemical kinase assays
A number of methods measure kinase activity by the detection of have used only the catalytic domain for HTS, ATP site–directed inhibitors
phosphorylated peptide. Several formats use micron-sized beads or are the primary class of agents identified. However, with a TR-FRET–
particles to partition the product phosphopeptide from ATP substrate based HTS of Akt-family kinases, Barnett et al.59 discovered a pharmaco-
or peptide. Radiometric filter binding remains the standard for kinase logically unique class of compounds using the full-length kinases. These
selectivity studies. For HTS, biotinylated 33P-phosphopeptide is typi- compounds stabilized the inactive conformation of Akt1 and Akt2 via a
cally captured by streptavidin-coated SPA beads41 or scintillant-treated site formed only in the presence of the pleckstrin homology (PH) domain
microtiter plates (for example, FlashPlates). In immobilized metal-ion of the kinase59, thereby allowing a unique mode of selectivity between
affinity-based fluorescence polarization (IMAP), transition metal-che- Akt1 and Akt2 versus Akt3, which is not inhibited by the chemical class
lated particles bind fluorescently labeled phosphopeptides with high owing to its lack of a PH domain.
affinity, thereby enabling an antibody-independent FP format for the
detection of phosphopeptide products52. Recently IMAP has been Proteases. As well-established drug targets60, proteases have received
adapted to FRET-based systems in which size limitations on the poly- considerable coverage in terms of assay formats and reagent kits.
peptide substrate (typically <10,000 MW) are relaxed53. A similar system Profluorescent or FRET-based substrates are commonly used to assay
has been developed based on fluorescence quenching through an iron- proteases as purified proteins or in cell lysates. Typically, probes are
phosphate interaction54. However, for HTS the ratiometric nature of coupled to a two-to-five-amino-acid peptide comprising a protease
IMAP is an advantage of this system over fluorescence quenching. cleavage site and become fluorescent after liberation by protease cleav-
Antibody-based formats that detect a specific phosphorylated peptide age13,61. Though ideally suited to HTS, a shortcoming of this approach
or sequence capitalize on the many phosphospecific antibodies available is that the fluorogenic peptide is often insufficient in length to cover the
and on nonseparation technologies, including TR-FRET, ALPHA and entire binding pocket.

a c
Gαi-coupled response Gαs-coupled response
EFC

100 100 100

EA
Activity (%)

Activity (%)

75 75 75
Activity (%)

ED

50 50 50

25 25 25

S
hν 0 0 0
Log[compound] (M) Log[compound] (M) Log[compound] (M)
TR-FRET b
hν 100 100 100
D A
Activity (%)

Activity (%)
Activity (%)

75 75 75

50 50 50

D 25 25 25
A
0 0 0
Log[compound] (M) Log[compound] (M) Log[compound] (M)

ED-labeled cAMP Acceptor-labeled cAMP Cellular cAMP


GPCR Gi AC [cAMP] AC Gs GPCR

Figure 2 Influence of assay design on the measured biological response. (a) An EFC-based competitive immunoassay illustrates an output signal that is
directly proportional to intracellular [cAMP]. (b) A TR-FRET output signal is indirectly proportional to intracellular [cAMP]. (c) The resulting responses for
hypothetical assay of Gi-coupled (left) and Gs-coupled (right) GPCRs are shown for the EFC (top) and TR-FRET assays (bottom) for compounds acting as
either agonists (dotted lines) or antagonists (solid lines). Assays having opposing signal outputs in response to the same analyte; here [cAMP] is an example
of an assay pair in which one could function in HTS and the other in folow-up confirmation and false positive detection.

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 469


REVIEW

Table 2 Equations for determining assay performance or sensitivity


Parameter Equation Comment
Coefficient of variation σ × 100 A measure of the precision relative to the mean value, calculated for the maximum and
%CV = —
µ minimum signals. An acceptable limit is <15%.
Signal to noise µmax – µmin A measure of the signal strength. This equation is sometimes given asa: µmax – µmin
S:N =
σmin σmax2 + σmin2
Signal to background µmax Usually calculated using control compounds, acceptable value >2-fold.
S:B =
µmin

Signal window µmax – µmin – 3(σmax + σmin) The significant signal between max and min controls, acceptable value >2-fold b.
SW =
σmax

Z' factor (3σmax + 3σmin) Representation of the SW using a score where a value >0.5 represents an acceptable
Z' = 1– assay. Z' is measured in the absence of library compounds; Z is in the presence.
µmax – µmin

Minimum significance ratio Smallest potency ratio between two measurements that is statistically significant
MSR = 102σd
(with 95% confidence).
Cheng-Prusoff IC50 Used to derive the Ki from the IC50 under conditions of competitive inhibition.
Ki =
(1 + [S]/Km)

Cheng-Prusoff IC50 Used to derive the Ki from the IC50 under conditions of competitive binding.
(ligand-binding) Ki =
(1 + [L]/Kd)

σ, s.d. of the assay signal; µ, mean of the assay signal; σd, s.d. of the difference in log potency; max, maximum signal; min, minimum signal. aThis is a redundant description of
the signal window equation. bSome assays, such as those with ratiometric data, may perform adequately with as little as a 20% change in signal (for example, see ref. 111) or with
0 < Z ′ < 0.5.

Although there are numerous ways to measure endoprotease activity, Using GFP fusions of NRs8, cell imaging assays have been described
assays for exoproteases that recognize C- or N-terminal residues are far that measure ligand-dependent, cytoplasm-to-nucleus translocation
less available62. The hydrolytic activity of the protease has been used of the glucocorticoid receptor (GR)72 or engineered NR chimeras73.
to process simple chromogenic esters and amides, but assays based on Translocation assays using GFP fusions of GR have been applied to
these substrates generally lack sensitivity for miniaturized HTS assays. 1,536-well systems (see ref. 8; PubChem BioAssay ID (AID) 450).
One approach is to use protease-mediated “epitope unmasking,” in Additionally, EFC assay formats have been described to measure NR
which cleavage by the exoprotease reveals a binding site for a detection translocation72,74 (PubChem AID 451). However, none of the aforemen-
reagent. Here, sensitive time-resolved FRET (TR-FRET)-based formats, tioned assays can distinguish the functional effect of active compounds
for example, can be applied63. on NR transactivation.
Cellular FRET-based methods have been developed using fluorescent To distinguish between agonist and antagonistic activity of com-
proteins or dyes linked by a protease consensus sequence. When expressed pounds, the targeting of NRs has been addressed with TR-FRET
or taken up by cells, cleavage of the probe can be measured by a change in ligand-dependent coregulator recruitment assays75, constructed with
fluorescence upon activation of proteases such as caspase-3 (ref. 64) and either coactivators or corepressors to identify agonists and antagonists.
hepatitis virus (HCV) NS2/3 (ref. 65). Similarly, cells expressing a fusion Additionally, cell-based reporter gene assays provide among the most
of GFP and alkaline phosphatase coupled through an NS3/4A cleavage site sensitive means to screen for functional activity. These assays fuse NR
were used to monitor the activity of this HCV protease by the detection response elements to reporters such as luciferase76, β-lactamase, and
of secreted alkaline phosphatase66. In an approach based on cell viability, secreted alkaline phosphatase (ref. 69 and other articles in the same
procaspase-3 was modified to contain a β-secretase cleavage motif that focus issue).
triggered apoptosis in 293T cells upon expression of β-secretase. β-secre-
tase inhibitors were shown in this assay to promote cell survival67. G protein–coupled receptors. GPCRs are targets of intensive drug
development by the pharmaceutical industry, and numerous, well-estab-
Nuclear receptors. NRs are transcription factors that are characterized lished HTS assay methods exist77,78 and continue to be developed79.
by common modular features including a DNA-binding domain (DBD) Technologies to measure ligand-receptor binding41,80 and the functional
and a ligand-binding domain (LBD). NRs are regulated by endobiotic consequence of receptor activation77 have enabled screens that target
ligands such as hormones and metabolites, or through xenobiotics68. the rich pharmacology of this receptor superfamily, including assays to
The human genome encodes ~48 NR members of known and unknown differentiate full, partial and inverse agonists, antagonists and allosteric
function (orphan NRs) that are actively being studied in drug discovery modulators81, and to identify ligands for orphan GPCRs8,82.
and basic research. Consequently, the ligand or DNA-binding properties GPCRs broadly orchestrate a wide range of cellular events and pro-
of isolated domains, and the cellular translocation and gene transactiva- cesses for which HTS assays have been configured, including guanine
tion properties of NRs, have been featured in assays designed for HTS nucleotide exchange5,83, modulation of second messenger concentra-
(ref. 69 and other articles in the same focus issue). tions84,85, ion channel activity86, protein phosphorylation56, protein
Classical competition binding assays for NRs can be divided into trafficking8,87,88, pigment dispersion89, gene transcription6 and cell
radiometric and fluorometric assays. SPA formats have been developed proliferation90, among others. The assays have different advantages and
for NRs such as peroxisome proliferator-activated receptor70, and FP has limitations, but collectively they illustrate the sophistication possible
been used for well-characterized receptors such as the steroidal NRs, in in HTS assay design given the current state of the art. For example,
which a fluorescent conjugate of a known receptor ligand and a purified GPCRs that activate intracellular Ca2+ stores can be assayed using cal-
LBD are prepared71. cium-sensitive dyes (for example, Fluo-3 and Fura-4) and rapid-inject

470 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY


REVIEW

imaging platforms84. However, GPCRs not naturally coupled to [Ca2+] ficking of ion channels represents another avenue for the development
modulation could not be assayed by this means until the development of HTS assays analogous to those developed for GPCRs. For example,
of so-called universal adaptors, molecular engineered G proteins91 or an ELISA-based assay has been reported that measures the amount of
promiscuous G proteins92, to reconfigure such GPCRs to stimulate Ca2+ cell surface hERG channel to determine drug effects on hERG channel
signaling. Today GPCR second messengers can be assayed by the direct trafficking101.
measurement of cAMP85 (Fig. 2) or inositol phosphate species57,93. Transporters, another category of transmembrane proteins, have an
Assays based on GPCR internalization are independent of G protein important role in ion, membrane potential and nutrient homeostasis.
subtype, unlike second messenger formation, and have been applied suc- Perturbations in transporter function can result in disease and altered
cessfully to analyze a number of GPCRs using either standard microtiter drug absorption, distribution, metabolism, excretion and toxicity
plate readers94 or fluorescence microscopy8,9. The use of the adaptor (ADMET) properties, leading to drug resistance, for example. As a case
protein β-arrestin 2, fused with GFP, permits an assay design that gen- in point, multidrug-resistance (MDR) transporters (for example, P-gly-
erally does not require modifications to the target receptor’s N or C coprotein) belong to the family of ATP-binding cassette (ABC) proteins
terminus. Such phenotypic assays enabled with automated high-content that extrude xenobiotic substances from the cell, including many drugs.
image analysis are bridging cell biology with HTS in a way that will have Assays based on ATPase activity and fluorophore transport have been
a major impact on chemical biology, as discussed below. described102,103. Typical assay formats for amino acid and neurotransmit-
ter transporters rely on radiolabeled substrates (for example, 3H-glycine)
Ion channels and transporters. Ion channels are membrane-spanning and involve multiple assay steps including medium aspirations. However,
proteins that regulate electrochemical gradient-driven movement of new assay methods are being developed that are suitable for HTS such as
inorganic ions such as Na+, K+, Ca2+ and Cl– into or out of cells. HTS yeast heterologous expression systems104 and anion sensors105.
technologies aimed at this target class are under rapid development95.
The existence of multiple responsive states, including closed, open and Assays for pathways and networks, and cellular phenotypes
inactive, to which drugs respond differently, add to the pharmacological The complexity of components and their interactions within cellular
diversity and assay development challenges96. networks offers a rich source of new targets to assay using technologies
Ligand binding, ion flux, and fluorescent probe-based assays are the to examine protein interactions, expression, distribution and stability.
established methods for ion channel HTS (Table 3 and Supplementary For instance, signaling mediated by Hedgehog, Wnt and Notch proteins
Table 1). Competition binding measures the affinity of the ligand- uses unconventional proteins implicated in developmental and disease
channel protein interaction but is dependent on the availability of processes, making these pathways important therapeutic targets106 that
radiolabeled ligands. As with GPCRs, binding assays do not detect the are assayable with HTS automated fluorescence microscopy methods8.
functional consequence of compound interactions with ion channels. The following sections describe additional cellular assays but reveal only
For these reasons, cell-based functional assays are more relevant to ion a glimpse of the breath of cellular processes amenable to testing by HTS
channel screening at both the primary and secondary levels. assays.
The calcium-sensing fluorophores Fluo-3, Fluo-4 and Fura-2 have
been used extensively in HTS for voltage- and ligand-gated channels Transcriptional readouts. Reporter gene expression is a mainstay for
and for GPCR-induced intracellular calcium release95. The fluorescence detection in cellular signaling assays6 and is useful for broadly identi-
intensity of these probes inside cells increases proportionally with the fying compounds that modulate the pathway and isolate new targets
elevation of intracellular free calcium concentration97. The addition within. However, the purposely promiscuous nature of these assays is
of opaque dyes to the assay buffer eliminates the need for cell washing disadvantageous if the objective is a probe of a particular component
steps, as only intracellular fluorescence of adherent cells is detected in of the pathway, as ‘target decryption’ can be a formidable undertaking.
bottom-read measurements, thereby improving compatibility with HTS A separate critical consideration resides in identifying a cell type that
protocols95. Voltage-sensing probes track changes in membrane poten- appropriately reflects the cellular context of the signaling pathway being
tial and have been used in several assay formats, such as FRET-based and assayed by the reporter; both the types and amounts of cell surface recep-
positional voltage sensors98. tors and intracellular signaling components can vary widely among the
Ion flux measurements using radioactive tracers such as 86Rb+, 45Ca2+, commonly used engineered cell lines.
22Na+ and 36Cl– are established assays of ion channel activity. However,

the medium removal steps and considerable radioactivity limit this Protein interactions and redistribution. Several innovative enzymatic
approach for HTS. Atomic absorbance spectrometry using nonradio- and image-based methodologies have provided new ways to identify
active tracer ions has recently permitted significant improvements to compounds that modulate movement and associations of protein com-
the ion flux assay for HTS99. ponents within living cells. These approaches fall into several general
Patch-clamp electrophysiology remains the standard for measuring categories: reporter complementation, resonance energy transfer and
ion channel activity, as the potency of certain compounds, especially protein redistribution. Low-affinity associations form the basis of pro-
those with state- or voltage-dependent mechanisms, may appear signifi- tein complementation assays (PCA), in which two proteins known to
cantly less potent when using other assay methods (for example, fluo- associate are separately fused to N- and C-terminal halves of a bifur-
rometric or rubidium ion flux). Automated patch-clamp instruments cated reporter protein. Upon interaction of the two sentinel proteins,
represent a remarkable development in technology, increasing through- the halves of the reporter (now in close proximity) reconstitute a func-
put from approximately ten compounds per week to hundreds95,100. tional enzymatic or fluorescent reporter. Commonly used PCA reporters
However, further improvements are needed in seal quality, throughput, include β-lactamase, GFP, YFP and luciferase27,40,107. This methodology
and cost before this technology can replace other formats such as fluo- is attractive because it can be adapted to many protein-protein interac-
rescence-based assays for HTS. tions and is amenable to screening and profiling, as demonstrated with
The broad examination of drug candidates for ion channel activity has YFP-based108 and β-lactamase109 reporters. Alternatively, high-affinity
become critical in the assessment of a candidate’s safety profile. In addi- β-galactosidase α complementation has been used to monitor protein
tion to the functional inhibition of ion channels, the folding and traf- movement. Here the ω fragment is targeted to a specific subcellular

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 471


REVIEW

Table 3 Assay technologies


Technology Description Refs.
AAS Atomic absorption spectroscopy for measurements of ion transport/flux (primarily focused on K + channels with Rb+ as tracer); 99
96-/384-well microtiter plates enabled with, for example, the ICR 12000.
Absorbance Follows Beer’s Law; trans- or epi-absorbance; via fluorescence attenuation. 42,43
ALPHA Amplified luminescent proximity homogeneous assay; donor-acceptor beads of ~200 nm diameter; λex 680 nm (laser); singlet 56,149
oxygen reacts with acceptor bead within ~200 nm distance; λem 520–620 nm.
Bioluminescence Renilla or firefly luciferases used in reporter gene assays; analogs of D-luciferin used to expand biochemical assays for 32
proteases, P450s, monoamine oxidases.
BRET Bioluminescence resonance energy transfer; useful for protein-protein interactions; Renilla luciferase donor luminescence (480 39,150
nm) stimulates fluorescence of GFP/YFP acceptor; follows dipole-dipole interaction 1/r 6 (~5 nm maximum).
Bla/CCF4 FRET-based assay using β-lactamase and β-lactam-containing substrate (for example, CCF4); λex 395 nm, λem 460/530 nm; 36,38,151
ratiometric reporter gene assays and other formats.
DELFIA Dissociation-enhanced lanthanide fluorescent immunoassay/time-resolved fluorescence; ELISA-like format using lanthanide 25
chelate-labeled antibody; τf ~500 µs; separation-based enhanced fluorescence; variations using labeled ligands; extremely
sensitive.
ECL Electrochemiluminescence assay. Uses microcarbon electrode plates allowing electrical oxidation of a [Ru(phen) 3]2+ label. 17
Multiplex applications in ELISA format.
EFC Enzyme fragment complementation. Based on β-galactosidase α complementation; high-affinity complementation. Applied to 57,74
both cell-free and cell-based systems.
FLT Fluorescent lifetime; time-domain method (~ns) measures the rate of fluorescent decay following excitation; promises to 146
eliminate fluorescence compound interference; only recently applied to some microplate systems.
FP Fluorescence polarization; polarized light reports change in probe ligand τc; fluorophores of τf ~ 5 ns needed to measure ligand 23,24
of MW < 1,500 binding to a receptor of MW ≥ 10,000; widely applied, ratiometric measurement.
FRET Fluorescence resonance energy transfer; donor-acceptor pair as in EDANS-DABCYL peptides or fluorescent proteins (for 22,53
example, CFP/YFP). Signal distance dependence ~5 nm maximum.
Fluorescence Measurement of fluorescence intensity including the use of chromophores for quenching as in profluorescent protease 22,152
substrates (~5 nm maximum distance); widely applied; superquenching varies in that nonequimolar amounts of donor
fluorophore and quencher are used. Includes kinetic modes of detection with timescales of approximately minutes to hours.
Fluorescence Requires fluorescent imaging plate reader or equivalent (for example, FLIPR, FDS6000, VIPR); timescale ~seconds; GPCR 84,98
(rapid kinetic) functional assays using Ca2+-sensitive dyes, or steady state ion channel activity using voltage-sensitive dyes.
GFP and variants Cell-based imaging microscopy/cytometry assays; translocation, redistribution, protein-protein interactions, reporter gene 27,112
assays; see below.
HT electrophysiology High-throughput electrophysiology; measurement of ion channel currents with microaperture arrays such as the planar 98,99,153
patch-clamp electrode (for example, PatchXpress 7000A, IonWorks Quattro).
PCA Protein complementation assay; recombinant split reporters (for example, GFP, luciferase, β-lactamase) fused to interacting 40,108
proteins; phenotypic assays; low-affinity complementation.
SPA, FlashPlate Scintillation proximity assay; nonseparation-based radiometric detection using solid supports; typical isotopes include 35S, 33P 41
or 3H.
TR-FRET Time-resolved FRET; also known commercially as HTRF/LANCE; nonseparation-based format using a lanthanide cryptate donor 25,26,53
fluorophore of long fluorescence lifetime (τf ~ 500 µs); detection distance ~9 nm max.
τf, fluorescent lifetime; τc, rotational correlation time.

region and the α-peptide-tagged translocating protein restores β-galac- ment of technology platforms for automated cellular imaging, together
tosidase activity once α and ω combine74. with sophisticated object recognition algorithms, has made the screen-
In the resonance energy transfer methods FRET and BRET (Table 3 ing of cell-based protein redistribution assays routine8,9. For instance,
and Supplementary Table 1), donor and acceptor (or sensor) proteins assays monitoring the redistribution of Akt1 (ref. 112), MAPK2 (ref.
are fused to the target proteins of interest. In BRET, a signal is generated 113) and ERF1 (ref. 114) have been used for HTS. Indeed, MacDonald
when a bioluminescent donor enzyme is brought into proximity with et al.108 report the use of 49 PCAs to monitor protein redistribution or
an acceptor fluorophore, the photon emission of which is measured39. interactions for profiling small-molecule activities.
In FRET, energy is transferred from a donor to an acceptor fluorophore
when the two proteins come into close proximity, which is manifested Protein stability. The addition of degradation and destabilization
as a change in wavelength of the emitted fluorescence. FRET-based domains to reporter proteins has been useful for assaying a number
sensors use spectrally distinct fluorescent proteins such as engineered of cell processes (such as ubiquitin-mediated degradation and the cell
GFP and YFP (reviewed in ref. 27). Intracellular indicators of specific cycle) and enabling screening of unstable proteins such as the cis-act-
kinase activity have been described using phosphorylation-activated ing viral NS2/3 protease65. For pathways in which a particular protein
FRET sensors engineered from fluorescent protein pairs110, or for is controlled through degradation, destabilization domain GFP sensors
insulin signaling via PIP3 recruitment of labeled Akt2 PH domains111. allow a measure of pathway activity. For instance, fusing the degrada-
Additionally, a cell-based HTS using BRET to detect CCR5-β arrestin tion and localization domains of cyclin B1 onto GFP conferred onto the
association identified inhibitors of CCR5-mediated signaling94. sensor the equivalent turnover and localization as endogenous cyclin
FRET, BRET, PCA and similar methods have been combined with B1, thereby allowing high-content screening for cell cycle modulators8.
subcellular imaging to produce highly successful and versatile HTS Similarly, fusion of the IκBα destabilization domain to luciferase was
assays for monitoring protein movements within cells9,110. The develop- used to identify modulators of NFκB signaling15. Also, coupling of

472 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY


REVIEW

the degradation domain of ornithine decarboxylase to the fluorescent a careful analysis of the screening results ensues to identify compounds
protein ZsGreen was used to discover new proteosome inhibitors115. with reliable biological activity, many times with lower (µM) potency.
Fusing only the degradation domain of a particular protein, as opposed Thus, the need for assay sensitivity is critically important to identify
to the entire protein, to a sensor is advantageous in that it minimizes compounds with half-maximal inhibitory concentrations (IC50s) in
the pathway interference that can be caused by overexpression of the the vicinity of the screening concentration, as these are false-negative
full-length protein8. candidates in assays having poor sensitivity.
Ideally, HTS assays are designed with substrate (S) or ligand (L) con-
Assays for rare disease targets. With over 1,800 genetic associations with centrations below their Km or Kd, so that the observed potency (IC50)
human disease now known116, the number of human diseases for which approximates the intrinsic potency Ki (see equations in Table 2; this
molecularly based assays could be developed is large. Occasionally a relationship differs for more complex mechanisms or when determin-
rare disease is caused by mutation of an immediately druggable tar- ing functional activity of receptors such as Kb values127–129). Though
get, and even more rarely one that may be ameliorated by a clinically increasing either [S] or [L] improves the overall Z' factor of the assay,
approved drug117. More commonly however, the genes responsible for doing so could potentially increase the IC50 values above the threshold
rare diseases have obscure or entirely unknown functions. Some com- used to identify actives in the screen (Fig. 3b,c), thereby preventing
mon causes of human genetic diseases are listed below and illustrate them from being identified as positives. Catalytic systems (for example,
the challenge of assay development for HTS in this arena: enzyme assays) should be designed for the minimum substrate con-
(i) Nonsense mutations resulting in premature termination and version needed to achieve an acceptable Z' factor. Although shifts in
absence of protein expression (for example, phenylketonuria, spinal IC50s can be less than two-fold for enzyme assays operating at conver-
muscular atrophy (SMA) and cystic fibrosis): assays to identify com- sions upward of 80%130, high turnover significantly affects the ability
pounds that compensate for the defect, for example by upregulating to detect inhibitors with potencies near the screening concentration
an alternative pathway or pseudogene as in SMA118, or promoting (Fig. 3d–f) and lowers the assay sensitivity for enzyme activators.
readthrough of premature stop codons119. Many HTS technologies are inherently unable to detect low-affin-
(ii) Missense mutations resulting in a potentially functional but mis- ity complexes (Kd ~ 1 µM) without resorting to high ligand or tar-
folded or mistrafficked protein (for example, cystic fibrosis): assay needs get concentration, or engineering in a high-affinity interaction. The
to discover compounds that act as molecular chaperones to facilitate various assays developed to screen for inhibitors of the protein-protein
correct folding and trafficking of the protein to its proper location120. interaction between the p53 tumor suppressor protein and its negative
(iii) Synonymous mutation that introduces a new splice donor or regulator (DM2) illustrate these issues (see Supplementary Table 2
acceptor site, thereby altering splicing and producing an abnormal online for assay details). In one assay the measurement of a p53-DM2
protein (for example, Hutchinson-Gilford Progeria Syndrome and interaction with a competition-binding FP assay131 required 1 µM DM2
β-thalassemia): assays to identify compounds that alter splicing121. to allow sufficient complex formation (~50% bound) for a detectable
(iv) Motif expansion or duplication mutations that result in expres- signal. This configuration limits the observable compound potency
sion of a protein with toxic or dominant negative function (for example, range, thereby rendering the assay unreceptive to weak compounds and
Huntington’s disease): assay needs to identify molecules that downregu- making it difficult to determine the rank order of potent compounds.
late the expression of the toxic protein or interfere with its function122. Conversely, high-affinity interactions resulting from enhanced avid-
These urgent and challenging biological problems will be an impor- ity can increase the apparent affinity of the binding partners many fold
tant frontier for HTS assay development in the next decade. through multiple receptor-ligand interactions. This “Velcro effect” is
evident as a rightward shift in IC50 values of competing control ligands.
Assay optimization for HTS For example, returning to the p53-DM2 interaction, Kane et al.132
A balance between statistical performance and assay sensitivity facili- developed a TR-FRET–based assay for this protein-protein interac-
tates the identification of the broadest range of compound potencies tion that required only nanomolar protein concentrations, in contrast
and efficacies from HTS. Assays are generally configured to measure one to the micromolar level needed in the previous FP format. However,
or more of the following changes: (i) concentration of a substance (for enhanced affinity due to multivalent interactions in the resulting com-
example, enzyme substrate or product, second messenger or reporter plex (Supplementary Table 2) decreased the sensitivity of the assay as
gene product); (ii) concentration of a receptor-ligand complex or (iii) judged by the higher-than-expected IC50 of unlabeled competitor p53
distribution of cellular markers. In addition, different methodologies protein. The potential for such avidity effects through surreptitious poly-
measuring the same biological process can have discordant determi- valent interactions should always be considered carefully when examin-
nations of apparent compound potency and efficacy37,123–125, which ing the benefits and limitations of the assay133. Further, irrespective of
emphasizes the need for selecting the proper assay format. affinity, assay conditions requiring a high fraction of bound ligand (for
In general, four types of assays can be defined based on the acceptable example, most FP-based assays) cannot rely on the Cheng-Prusoff cor-
limits of Z' factor and assay sensitivity (Fig. 3a). Conditions leading rection to estimate the Kd or Ki of control or test compounds. Rather,
to satisfactory performance (Z' factor > 0.5) should be identified that accurate approximations can be made using equations describing the
minimally impact assay sensitivity of a control ligand or stimulus (Fig. complete equilibria of the competing components134. In another bio-
3a, quadrant i). With this in mind, the optimal assay will not always chemical format, the thermal shift assay135, which measures a ligand’s
have the highest Z' factor, as some very robust assays (Z' scores > 0.8) ability to stabilize a protein’s thermal denaturation, has been used to
can have poor sensitivity (Fig. 3a, quadrant iii). Conversely, assays with investigate compound binding to the p53 binding domain of DM2
lower Z' factors may have superior sensitivity (for example, Fig. 3a, directly. Though thermal shift technology at present is relatively low-
quadrants i and ii) and may be rescued for HTS using replicate or throughput, requiring comparatively large amounts of target protein,
concentration-dependent formats126. Because potent compounds can it is one of the few function- or ligand-independent methods to screen
be identified in poorly designed assays, the irrational exuberance of for small molecule–protein interactions that may provide a means to
observing such high-potency or high-efficacy compounds is often dis- identify small molecules with protein complex disruption potential
placed by the realization that the activity was artifactual in nature, and (Supplementary Tables 1 and 2).

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 473


REVIEW

a b c 12 1.0
1
100
10
0.8
i iii

Fold (IC50/Ki)
75 8 Z' factor

Inhibition (%)

Z' factor
0.6
Z' factor

0.5 6
50
0.4
4
ii iv
25 Fold
0.2
2

0 0.0
0 0
3 10 –6 –5 –4 0.2 0.4 0.6 0.8 1.0

Fold (IC50´/IC50) Log[inhibitor] (M) Fraction bound receptor

d e f
12 1.0 12 1.0
100
10 10
0.8 0.8
Z' factor Z' factor

Fold (IC50´/IC50)
Fold (IC50´/IC50)

75 8 8
Inhibition (%)

Z' factor
0.6 0.6

Z' factor
6 6
50
0.4 0.4
4 4
Fold Fold
25 0.2 0.2
2 2

0 0 0.0 0 0.0
–9 –8 –7 –6 –5 –4 0 25 50 75 100 0 25 50 75 100

Log[inhibitor] (M) Conversion (%) Conversion (%)


Figure 3 Assay performance and sensitivity. (a) Assay performance (Z' factor) is plotted against sensitivity, as defined by the ratio of observed (IC50’)
and known (IC50) values of a control compound. i, Optimal performance and sensitivity (Z' > 0.5, IC50’/IC50 < three-fold). ii, Poor performance and good
sensitivity (Z' < 0.5). iii, Good performance and poor sensitivity (Z' > 0.5; IC50’/IC50 > three-fold). iv, Poor performance and sensitivity (Z' < 0.5; IC50’/IC50
> three-fold). (b) Effect of varying the [ligand]/Kd ratio for a binding assay. Concentration-response (CR) curves represent [ligand]/Kd ratios as follows: 0.10,
blue; 1, green; 3, orange; 5, black; and 11, purple. IC50’ was calculated using the Cheng-Prusoff equation (Table 2), with a compound Ki = 5 µM and a
ligand Kd = 10 µM. The gray box is the “observation window” representing observed activity for a 10 µM screening concentration and a hit cutoff of >30%
inhibition. (c) A modeled assay in which between ~30% and 70% bound receptor yields optimal performance. (d) CR curves for an enzyme assay at different
substrate conversions (10%, red; 40%, blue; 60%, green; 80%, orange; 90%, black and 98%, purple) for two reversible inhibitors (solid lines, IC 50 = 100
nM; dotted lines, IC50 = 10 µM). The observation window (gray box) indicates how shifts in IC50’ limit the identification of weak inhibitors, including potent,
low-efficacy compounds (gray dotted curve). (e,f) Shifts in IC50’ were calculated as described130. Two modeled enzyme assays in which either an adequate
performance (for example, two-fold S:B and %CV < 5%) is achieved at 10% conversion (e) or the equivalent performance is not reached until ~60%
substrate conversion (f). Grey regions indicate areas of either poor assay performance or poor assay sensitivity (a, c, e and f).

Owing to recent advances in high-throughput cellular imaging9,136, tration to achieve a suitable signal139. The net result of increasing target
phenotypic assays now represent a viable alternative or complement to concentration is to increase assay cost and decrease sensitivity140.
biochemical formats for protein-protein interaction HTS assays. Several
cell-based assays have been used to investigate the p53-DM2 interac- Compound interference within HTS assays
tion108,137, and these may circumvent the affinity and ‘enhanced avidity’ Compound-dependent assay interference is a very common cause
concerns encountered in systems configured with isolated proteins. of screening artifacts and must be ruled out before any compound
Reagent-coated surfaces provide both opportunities and drawbacks activity is attributed to modulating the biological activity of interest.
for HTS assays. Formats involving reagent-coated beads or surfaces Here the inhibition or activation of an assay signal is not caused by the
enable a physical separation of a bound and free ligand, thereby elimi- compound’s effect on the target or pathway of interest, but rather it
nating filtration steps41,133, or they provide components that comprise is due to the compound’s ability to mimic that effect as a result of its
the detection system, as in ALPHA138. However, two issues should be physicochemical properties (for example, light absorbance) or adventi-
considered when designing heterogeneous assays using solid supports. tious biological activities (for example, luciferase inhibition)141,142 (Fig.
One issue pertains to the “avidity” of the interaction mentioned above 4). Cheminformatic approaches have been described that attempt to
that can occur in bead-based assays in which two binding partners are identify or remove from screening collections compounds with such
coated or tethered to separate beads, which results in multiple binding problematic characteristics143, but currently prediction of such prop-
sites on the solid support (for example, the Velcro effect). The other erties is not reliable, and each active compound must be assumed to
issue is the capacity of the solid support to ‘hold’ a reagent or substrate. be an artifact until proven otherwise. This may be particularly true
For example, when presenting an enzyme substrate on a solid support for chemical genomics applications, as the allowed functionalities of a
the amount of substrate that can be coated onto a surface can limit the chemical probe are less strict than for a medicinal lead. Thus although
reaction rate, and is often compensated by increasing enzyme concen- the boundaries of a useful chemical series may be broader for chemical

474 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY


REVIEW

genomics applications than for drug development, the potential for signal strength or brightness and can determine assay sensitivity to
misleading HTS activity is perhaps even greater with ‘probe-like’ than certain classes of interferences, such as attenuation of luminescence
with ‘drug-like’ compounds. or scintillation by light-absorbing compounds142. Signal brightness
Compound fluorescence and absorbance are very common causes becomes particularly important with regard to compound interfer-
of signal amplification or attenuation. Heterocyclic compounds that ence when incident light is used to excite a fluorophore. In this case,
dominate screening collections most commonly fluoresce in the assays having emitted light signals that overlap with the fluorescence
blue-green range, meaning that common assay sensors, such as GFP from compounds will show increased false positives (Fig. 4a,b). The
and coumarin, as well as the many profluorescent enzyme substrates amount of compound fluorescence is dependent on the wavelength
that excite or emit near the blue end of the spectrum, are susceptible of excitation energy used, with blue light being more problematic
to this type of optical interference. Assay responses can differ in than red light for typical heterocyclic compound libraries142,144.

a b c

Excitation light

hν hν

Emission
light

Positives False positives

Figure 4 Compound effects on assay outputs. (a–c) The effect of different types of compound interference within a library (colored slices of the pie charts) on
the scoring of actives. Interfering compounds are represented as dark spots arranged as a shadow cast by assay output signal, and individual spots represent
genuine actives in assays using different modes of detection (blue or red fluorescence, or luminescence). In assays in which blue spectrum excitation is
used (a), significant interference from compound emission is possible, whereas for red-shifted excitation light (b), interference is typically reduced. Also
shown below are example fluorescent compounds 5 and 6, excited by blue and red light, respectively. (c) Luminescence assays do not use an external
excitation source, which results in negligible interference by compound fluorescence unless compounds or sample impurities absorb light (for example, inner
filter effect by 6). (d) Example molecules that illustrate different interference mechanisms. Biotin (1) analogs such as 1a disrupt biotin-avidin complexes
commonly used in assay designs (for example, PubChem AIDs 595 and 632). The imidazole (2) derivative 2a can coordinate with Ni2+ and interfere with
polyhistidine-tagged proteins at µM concentrations. Nickel-chelated microspheres are used in some instances to tether polyhistidine-tagged proteins (for
example, PubChem AIDs 595 and 632), and can thus be disrupted by compounds similar to 2a at typical screening concentrations. The luciferin (3) analog
3a is an inhibitor of the reporter luciferase. A major source of interference is compound aggregation (see 4), in which detergent-sensitive ‘promiscuous’
inhibitors block assays by forming target-sequestering/denaturing compound aggregates (4a)147.

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 475


REVIEW

To provide experimental data on the fluorescent properties of the HTS assay. Assay sensitivity and susceptibility to compound interference
compounds screened in our center, we recently performed spectro- must also be studied prospectively when designing assays for HTS and
scopic profiling of ~70,000 compounds from the Molecular Libraries planning follow-up analysis of active samples.
Small Molecule Repository for eight common wavelengths (unpub- In general, the signal output of an assay should be assessed in con-
lished data; PubChem AIDs 587–592). For example, 4-methylumbel- junction with consequential controls. For example, protein-protein
liferone (4-MU; excitation 385 nm, emission 502 nm) is a commonly interaction assays using purified proteins may have suitable statistical
used label in profluorescent substrates for a wide range of hydrolytic performance factors but show poor sensitivity, which suggests that cel-
enzymes. We found ~2% of the library fluoresced at a level equivalent to lular assays with phenotypic outputs should be considered as alterna-
at least 100 nM 4-MU, which indicates that a screen using this excitation tives. When investigating novel target biology, effort should be made
wavelength would yield many false positives via this mechanism. to leverage existing well-developed technologies where possible. For
Several approaches can minimize interference from fluorescent instance, methods developed to detect ATP and ADP in protein kinase
compounds, and should be used whenever possible. Separation-based assays48,51,53 are also useful for lipid49 and metabolic kinases126.
formats such as ELISA offer an advantage over solution-phase assays, As HTS assay technologies, screening systems, and analytical instru-
as compounds are removed before the detection step. However, the mentation such as patch clamping and mass spectrometry evolve, the
necessary separation steps can make ELISA difficult to miniaturize. For interfacing of large compound libraries with sophisticated assay and
homogeneous assays, compound fluorescence can be reduced by using detection platforms will greatly expand the capability to identify chemi-
probes that use red-shifted fluorophores61, time-resolved lanthanide- cal probes for the vast untapped biology encoded by genomes. The rami-
based fluorescence25,26, pre-reads or multiple time points that permit fications will allow a reinterpretation of the ‘druggable genome’.
background subtraction or rate determinations145. Technologies that
Note: Supplementary information is available on the Nature Chemical Biology website.
in principle are minimally affected by compound fluorescence, such as
fluorescence lifetime (FLT) spectroscopy146, are in development, but for ACKNOWLEDGMENTS
now remain unproven in HTS. We thank D. Leja and J. Fekces for illustrations. The authors are supported by the
Signal attenuation or activation of assay signal can also be caused Molecular Libraries Initiative of the NIH Roadmap for Medical Research and the
Intramural Research Program of the National Human Genome Research Institute,
by compound aggregation, which can result in target denaturation,
NIH.
ligand sequestration or light scattering147 (Fig. 4c,d). Like the optical
properties discussed above, aggregation often occurs within the concen- COMPETING INTERESTS STATEMENT
tration range of the expected biological response and so may be easily The authors declare no competing financial interests.
mistaken for true modulation of the target biology. Aggregation effects Published online at http://www.nature.com/naturechemicalbiology
are most prevalent at concentrations >1 µM, the hallmark of which is Reprints and permissions information is available online at http://npg.nature.com/
the elimination of apparent compound activity by the addition of deter- reprintsandpermissions
gents such as Triton X-100, a disruptor of aggregate micellar structures.
Such compounds may show steep Hill slopes, but this alone does not
1. Zhang, R. et al. A continuous spectrophotometric assay for the hepatitis C virus
distinguish aggregation from biologically relevant, stoichiometric, or serine protease. Anal. Biochem. 270, 268–275 (1999).
cooperative modulation of enzyme activity. The prevalence of aggrega- 2. McDonald, O.B. et al. A scintillation proximity assay for the Raf/MEK/ERK kinase
cascade: high-throughput screening and identification of selective enzyme inhibi-
tion-dependent enzyme inhibition has been observed to range from 2 tors. Anal. Biochem. 268, 318–329 (1999).
to 10%, depending on the concentration of detergent and the chemi- 3. Seville, M., West, A.B., Cull, M.G. & McHenry, C.S. Fluorometric assay for DNA
cal library tested147; importantly, no clear structure-activity relation- polymerases and reverse transcriptase. Biotechniques 21, 664–672 (1996).
4. Verma, R. et al. Ubistatins inhibit proteasome-dependent degradation by binding
ship has been evident from these studies, making the establishment of the ubiquitin chain. Science 306, 117–120 (2004).
aggregation behavior currently entirely empirical. Results of a large 5. Ferrer, M. et al. A fully automated [35S]GTPgammaS scintillation proximity assay
aggregation screen recently performed at our center can be found in for the high-throughput screening of Gi-linked G protein-coupled receptors. Assay
Drug Dev. Technol. 1, 261–273 (2003).
PubChem (AIDs 585 and 584). 6. Dinger, M.C. & Beck-Sickinger, A.G. in Molecular Biology in Medicinal Chemistry
Biological interferences are the second major source of compound- (eds. Dingermann, T., Steinhiber, D. & Folkers, G.) 73–94 (Wiley-VCH Verlag
GmbH & Co., Weinheim, Germany, 2004).
specific assay artifacts (Fig. 4a). These are due to inhibition (or, less
7. Hodder, P., Mull, R., Cassaday, J., Berry, K. & Strulovici, B. Miniaturization of
frequently, activation) of a detection component of the assay such as intracellular calcium functional assays to 1536-well plate format using a fluoro-
inhibitors of enzyme reporters (for example, luciferases or β-lactamases) metric imaging plate reader. J. Biomol. Screen. 9, 417–426 (2004).
8. Inglese, J. (ed.) Measuring Biological Responses with Automated Microscopy Vol.
or intermolecular interactions (for example, nickel chelate–polyhisti- 414 (Elsevier Academic, San Diego, 2006).
dine tag, small-molecule antigen-antibody, biotin-streptavidin). As an 9. Taylor, D.L., Haskins, J.R. & Giuliano, K.A. (eds.) High Content Screening Vol.
example, inhibition of luciferase by library compounds is common and 356 (Humana, Totowa, New Jersey, 2006).
10. Sever, J.L. Application of a microtechnique to viral serological investigations.
has been described for the well-publicized compound resveratrol148. J. Immunol. 88, 320–329 (1962).
A recent screen of P. pyrali luciferase showed between 2 and 3% of the 11. Carroll, S.S., Inglese, J., Mao, S.-S. & Olsen, D.B. in Molecular Cancer
library was capable of interfering with luciferase detection (unpub- Therapeutics: Strategies for Drug Discovery and Development (ed. Prendergast,
G.C.) 119–140 (John Wiley & Sons, Inc., Hoboken, New Jersey, USA, 2004).
lished data; PubChem AID 411). 12. Wölcke, J. & Ullmann, D. Miniaturized HTS technologies - uHTS. Drug Discov.
Today 6, 637–646 (2001).
13. Karvinen, J. et al. Caspase multiplexing: simultaneous homogeneous time-
Summary resolved quenching assay (TruPoint) for caspases 1, 3, and 6. Anal. Biochem.
Given the right compound library, assay design becomes the determining 325, 317–325 (2004).
factor for identifying chemical probes using HTS. Before the selection 14. Bandyopadhyay, S. et al. A high-throughput drug screen targeted to the 5′untrans-
lated region of Alzheimer amyloid precursor protein mRNA. J. Biomol. Screen.
of an assay format for HTS, thorough consideration of its advantages, 11, 469–480 (2006).
limitations, and (when possible) prior application in HTS is recom- 15. Davis, R.E. et al. A cell-based assay for IkappaBalpha stabilization using a
mended. The investigator’s familiarity with specific technologies should two-color dual luciferase-based sensor. Assay Drug Dev. Technol. 5, 85–104
(2007).
not form the basis of an assay strategy; rather, an unbiased assessment 16. O’Boyle, D.R. II et al. Development of a cell-based high-throughput specificity
of key factors discussed here will increase the probability of a successful screen using a hepatitis C virus-bovine viral diarrhea virus dual replicon assay.

476 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY


REVIEW

Antimicrob. Agents Chemother. 49, 1346–1353 (2005). 51. Charter, N.W., Kauffman, L., Singh, R. & Eglen, R.M. A generic, homogenous
17. Ross, S. et al. High-content screening analysis of the p38 pathway: profiling of method for measuring kinase and inhibitor activity via adenosine 5′-diphosphate
structurally related p38alpha kinase inhibitors using cell-based assays. Assay accumulation. J. Biomol. Screen. 11, 390–399 (2006).
Drug Dev. Technol. 4, 397–409 (2006). 52. Sportsman, J.R., Gaudet, E.A. & Boge, A. Immobilized metal ion affinity-based
18. Iversen, P.W., Eastwood, B.J., Sittampalam, G.S. & Cox, K.L. A comparison of fluorescence polarization (IMAP): advances in kinase screening. Assay Drug Dev.
assay performance measures in screening assays: signal window, Z’ factor, and Technol. 2, 205–214 (2004).
assay variability ratio. J. Biomol. Screen. 11, 247–252 (2006). 53. Comley, J. Kinase screening and profiling-spoilt for choice. Drug Discov. World
19. Eastwood, B.J. et al. The minimum significant ratio: a statistical parameter to 7, 27–50 (2006).
characterize the reproducibility of potency estimates from concentration-response 54. Morgan, A.G., McCauley, T.J., Stanaitis, M.L., Mathrubutham, M. & Millis, S.Z.
assays and estimation by replicate-experiment studies. J. Biomol. Screen. 11, Development and validation of a fluorescence technology for both primary and
253–261 (2006). secondary screening of kinases that facilitates compound selectivity and site-spe-
20. Assay Guidance Manual Committee. Assay Guidance Manual Version 4.1 (Eli cific inhibitor determination. Assay Drug Dev. Technol. 2, 171–181 (2004).
Lilly and Company and NIH Chemical Genomics Center, Bethesda, Maryland, 55. Newbatt, Y. et al. Identification of inhibitors of the kinase activity of oncogenic
2005). V600E BRAF in an enzyme cascade high-throughput screen. J. Biomol. Screen.
21. Fluorescence Spectra Viewer (Invitrogen Corporation, Carlsbad, California, 2006). 11, 145–154 (2006).
<http://probes.invitrogen.com/servlets/spectraviewer?fileid1=1304dna> 56. Osmond, R.I. et al. GPCR screening via ERK 1/2: a novel platform for screening
22. Lakowicz, J.R. Principles of Fluorescence Spectroscopy (Springer, Berlin, G protein-coupled receptors. J. Biomol. Screen. 10, 730–737 (2005).
2006). 57. Eglen, R.M. & Singh, R. Beta galactosidase enzyme fragment complementation as
23. Owicki, J.C. Fluorescence polarization and anisotropy in high throughput screen- a novel technology for high throughput screening. Comb. Chem. High Throughput
ing: perspectives and primer. J. Biomol. Screen. 5, 297–306 (2000). Screen. 6, 381–387 (2003).
24. Jameson, D.M. & Croney, J.C. Fluorescence polarization: past, present and future. 58. Zaman, G.J., van der Lee, M.M., Kok, J.J., Nelissen, R.L. & Loomans, E.E.
Comb. Chem. High Throughput Screen. 6, 167–176 (2003). Enzyme fragment complementation binding assay for p38alpha mitogen-activated
25. Hemmila, I. & Laitala, V. Progress in lanthanides as luminescent probes. J. protein kinase to study the binding kinetics of enzyme inhibitors. Assay Drug Dev.
Fluoresc. 15, 529–542 (2005). Technol. 4, 411–420 (2006).
26. Trinquet, E. & Mathis, G. Fluorescence technologies for the investigation of 59. Barnett, S.F. et al. Identification and characterization of pleckstrin-homology-
chemical libraries. Mol. Biosyst. 2, 380–387 (2006). domain-dependent and isoenzyme-specific Akt inhibitors. Biochem. J. 385,
27. Giepmans, B.N., Adams, S.R., Ellisman, M.H. & Tsien, R.Y. The fluorescent 399–408 (2005).
toolbox for assessing protein location and function. Science 312, 217–224 60. Leung, D., Abbenante, G. & Fairlie, D.P. Protease inhibitors: current status and
(2006). future prospects. J. Med. Chem. 43, 305–341 (2000).
28. Haugland, R.P. The Handbook—A Guide to Fluorescent Probes and Labeling 61. Grant, S.K., Sklar, J.G. & Cummings, R.T. Development of novel assays for proteo-
Technologies 10th edn. (Eugene, Oregon, 2006). lytic enzymes using rhodamine-based fluorogenic substrates. J. Biomol. Screen.
29. Medintz, I.L., Uyeda, H.T., Goldman, E.R. & Mattoussi, H. Quantum dot biocon- 7, 531–540 (2002).
jugates for imaging, labelling and sensing. Nat. Mater. 4, 435–446 (2005). 62. Auld, D. in Proteolytic Enzymes Tools: and Targets (eds. Stocker, W. & Sterchi,
30. Haney, S.A., LaPan, P., Pan, J. & Zhang, J. High-content screening moves to the E. E.) 30–48 (Springer-Velag, Berlin Heidelberg, 1999).
front of the line. Drug Discov. Today 11, 889–894 (2006). 63. Ferrer, M. et al. Miniaturizable homogenous time-resolved fluorescence assay for
31. Bowen, W.P. & Wylie, P.G. Application of laser-scanning fluorescence microplate carboxypeptidase B activity. Anal. Biochem. 317, 94–98 (2003).
cytometry in high content screening. Assay Drug Dev. Technol. 4, 209–221 64. Jones, J., Heim, R., Hare, E., Stack, J. & Pollok, B.A. Development and applica-
(2006). tion of a GFP-FRET intracellular caspase assay for drug screening. J. Biomol.
32. Fan, F. & Wood, K.V. Bioluminescent assays for high-throughput screening. Assay Screen. 5, 307–317 (2000).
Drug Dev. Technol. 5, 127–136 (2007). 65. Whitney, M. et al. A collaborative screening program for the discovery of inhibi-
33. Nakatsu, T. et al. Structural basis for the spectral difference in luciferase biolu- tors of HCV NS2/3 cis-cleaving protease activity. J. Biomol. Screen. 7, 149–154
minescence. Nature 440, 372–376 (2006). (2002).
34. Blair, W.S. et al. A novel HIV-1 antiviral high throughput screening approach for 66. Lee, J.C. et al. Development of a cell-based assay for monitoring specific hepa-
the discovery of HIV-1 inhibitors. Antiviral Res. 65, 107–116 (2005). titis C virus NS3/4A protease activity in mammalian cells. Anal. Biochem. 316,
35. Hao, W. et al. Development of a novel dicistronic reporter-selectable hepatitis C 162–170 (2003).
virus replicon suitable for high-throughput inhibitor screening. Antimicrob. Agents 67. Li, Y., Liu, Y.W. & Cordell, B. Novel functional assay for proteases and modulators.
Chemother. 51, 95–102 (2007). Application in beta-secretase studies. Mol. Biotechnol. 18, 1–10 (2001).
36. Zlokarnik, G. Fusions to beta-lactamase as a reporter for gene expression in live 68. Olefsky, J.M. Nuclear receptor minireview series. J. Biol. Chem. 276, 36863–
mammalian cells. Methods Enzymol. 326, 221–244 (2000). 36864 (2001).
37. Kunapuli, P. et al. Development of an intact cell reporter gene beta-lactamase 69. Zeh, K. et al. Gain-of-function somatic cell lines for drug discovery applications
assay for G protein-coupled receptors for high-throughput screening. Anal. generated by homologous recombination. Assay Drug Dev. Technol. 1, 755–765
Biochem. 314, 16–29 (2003). (2003).
38. Oosterom, J., van Doornmalen, E.J., Lobregt, S., Blomenrohr, M. & Zaman, G.J. 70. Sun, S., Almaden, J., Carlson, T.J., Barker, J. & Gehring, M.R. Assay develop-
High-throughput screening using beta-lactamase reporter-gene technology for ment and data analysis of receptor-ligand binding based on scintillation proximity
identification of low-molecular-weight antagonists of the human gonadotropin assay. Metab. Eng. 7, 38–44 (2005).
releasing hormone receptor. Assay Drug Dev. Technol. 3, 143–154 (2005). 71. Parker, G.J., Law, T.L., Lenoch, F.J. & Bolger, R.E. Development of high through-
39. Pfleger, K.D. & Eidne, K.A. Illuminating insights into protein-protein interac- put screening assays using fluorescence polarization: nuclear receptor-ligand-
tions using bioluminescence resonance energy transfer (BRET). Nat. Methods binding and kinase/phosphatase assays. J. Biomol. Screen. 5, 77–88 (2000).
3, 165–174 (2006). 72. Fung, P. et al. A homogeneous cell-based assay to measure nuclear translocation
40. Remy, I. & Michnick, S.W. Mapping biochemical networks with protein-fragment using beta-galactosidase enzyme fragment complementation. Assay Drug Dev.
complementation assays. Methods Mol. Biol. 261, 411–426 (2004). Technol. 4, 263–272 (2006).
41. Wu, S. & Liu, B. Application of scintillation proximity assay in drug discovery. 73. Mackem, S., Baumann, C.T. & Hager, G.L. A glucocorticoid/retinoic acid receptor
BioDrugs 19, 383–392 (2005). chimera that displays cytoplasmic/nuclear translocation in response to retinoic
42. Lavery, P., Brown, M.J. & Pope, A.J. Simple absorbance-based assays for ultra- acid. A real time sensing assay for nuclear receptor ligands. J. Biol. Chem. 276,
high throughput screening. J. Biomol. Screen. 6, 3–9 (2001). 45501–45504 (2001).
43. Zuck, P. et al. Miniaturization of absorbance assays using the fluorescent proper- 74. Wehrman, T.S., Casipit, C.L., Gewertz, N.M. & Blau, H.M. Enzymatic detection
ties of white microplates. Anal. Biochem. 342, 254–259 (2005). of protein translocation. Nat. Methods 2, 521–527 (2005).
44. Zheng, C.J. et al. Therapeutic targets: progress of their exploration and investiga- 75. Gowda, K., Marks, B.D., Zielinski, T.K. & Ozers, M.S. Development of a coactiva-
tion of their characteristics. Pharmacol. Rev. 58, 259–279 (2006). tor displacement assay for the orphan receptor estrogen-related receptor-gamma
45. Fox, S. et al. High-throughput screening: update on practices and success. using time-resolved fluorescence resonance energy transfer. Anal. Biochem. 357,
J. Biomol. Screen. 11, 864–869 (2006). 105–115 (2006).
46. Hopkins, A.L. & Groom, C.R. The druggable genome. Nat. Rev. Drug Discov. 1, 76. Zhu, Z. et al. Correlation of high-throughput pregnane X receptor (PXR) transac-
727–730 (2002). tivation and binding assays. J. Biomol. Screen. 9, 533–540 (2004).
47. Pawson, T. & Scott, J.D. Protein phosphorylation in signaling—50 years and 77. Eglen, R.M., Bosse, R. & Reisine, T. Emerging concepts of guanine nucleo-
counting. Trends Biochem. Sci. 30, 286–290 (2005). tide-binding protein-coupled receptor (GPCR) function and implications for high
48. Singh, P., Harden, B.J., Lillywhite, B.J. & Broad, P.M. Identification of kinase throghput screening. Assay Drug Dev. Technol. 5, 425–451 (2007).
inhibitors by an ATP depletion method. Assay Drug Dev. Technol. 2, 161–169 78. Jacoby, E., Bouhelal, R., Gerspacher, M. & Seuwen, K. The 7 TM G-protein-
(2004). coupled receptor target family. ChemMedChem 1, 761–782 (2006).
49. Munagala, N. et al. Identification of small molecule ceramide kinase inhibitors 79. Yu, N. et al. Real-time monitoring of morphological changes in living cells by
using a homogeneous chemiluminescence high throughput assay. Assay Drug electronic cell sensor arrays: an approach to study G protein-coupled receptors.
Dev. Technol. 5, 65–73 (2007). Anal. Chem. 78, 35–43 (2006).
50. Lowery, R.G. & Kleman-Leyer, K. Transcreener: screening enzymes involved in 80. Handl, H.L. & Gillies, R.J. Lanthanide-based luminescent assays for ligand-recep-
covalent regulation. Expert Opin. Ther. Targets 10, 179–190 (2006). tor interactions. Life Sci. 77, 361–371 (2005).

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 477


REVIEW

81. Kenakin, T. Inverse, protean, and ligand-selective agonism: matters of receptor 113. Almholt, D.L. et al. Nuclear export inhibitors and kinase inhibitors identified
conformation. FASEB J. 15, 598–611 (2001). using a MAPK-activated protein kinase 2 redistribution screen. Assay Drug Dev.
82. Behan, D.P. & Chalmers, D.T. The use of constitutively active receptors for drug Technol. 2, 7–20 (2004).
discovery at the G protein-coupled receptor gene pool. Curr. Opin. Drug Discov. 114. Granas, C. et al. Identification of RAS-mitogen-activated protein kinase signaling
Devel. 4, 548–560 (2001). pathway modulators in an ERF1 redistribution screen. J. Biomol. Screen. 11,
83. Labrecque, J. et al. The development of an europium-GTP assay to quantitate che- 423–434 (2006).
mokine antagonist interactions for CXCR4 and CCR5. Assay Drug Dev. Technol. 115. Rickardson, L., Wickstrom, M., Larsson, R. & Lovborg, H. Image-based screen-
3, 637–648 (2005). ing for the identification of novel proteasome inhibitors. J. Biomol. Screen. 12,
84. Chambers, C. et al. Measuring intracellular calcium fluxes in high throughput 203–210 (2007).
mode. Comb. Chem. High Throughput Screen. 6, 355–362 (2003). 116. Online Mendelian Inheritance of Man (OMIM). (Johns Hopkins University and
85. Williams, C. cAMP detection methods in HTS: selecting the best from the rest. National Center for Biotechnology Information, National Library of Medicine,
Nat. Rev. Drug Discov. 3, 125–135 (2004). 2007). <http://www.ncbi.nlm.nih.gov/omim/>
86. Reinscheid, R.K., Kim, J., Zeng, J. & Civelli, O. High-throughput real-time moni- 117. Habashi, J.P. et al. Losartan, an AT1 antagonist, prevents aortic aneurysm in a
toring of Gs-coupled receptor activation in intact cells using cyclic nucleotide- mouse model of Marfan syndrome. Science 312, 117–121 (2006).
gated channels. Eur. J. Pharmacol. 478, 27–34 (2003). 118. Jarecki, J. et al. Diverse small-molecule modulators of SMN expression found
87. Adie, E.J. et al. CypHer 5: a generic approach for measuring the activation and by high-throughput compound screening: early leads towards a therapeutic for
trafficking of G protein-coupled receptors in live cells. Assay Drug Dev. Technol. spinal muscular atrophy. Hum. Mol. Genet. 14, 2003–2018 (2005).
1, 251–259 (2003). 119. Welch, E.M. et al. PTC124 targets genetic disorders caused by nonsense muta-
88. Charest, P.G., Terrillon, S. & Bouvier, M. Monitoring agonist-promoted confor- tions. Nature 447, 87–91 (2007).
mational changes of beta-arrestin in living cells by intramolecular BRET. EMBO 120. Bernier, V., Lagace, M., Bichet, D.G. & Bouvier, M. Pharmacological chaperones:
Rep. 6, 334–340 (2005). potential treatment for conformational diseases. Trends Endocrinol. Metab. 15,
89. Karlsson, A.M. et al. Biosensing of opioids using frog melanophores. Biosens. 222–228 (2004).
Bioelectron. 17, 331–335 (2002). 121. Sazani, P. et al. Nuclear antisense effects of neutral, anionic and cationic oligo-
90. Burstein, E.S. et al. Integrative functional assays, chemical genomics and high nucleotide analogs. Nucleic Acids Res. 29, 3965–3974 (2001).
throughput screening: harnessing signal transduction pathways to a common HTS 122. Zhang, X. et al. A potent small molecule inhibits polyglutamine aggregation in
readout. Curr. Pharm. Des. 12, 1717–1729 (2006). Huntington’s disease neurons and suppresses neurodegeneration in vivo. Proc.
91. Liu, A.M. et al. Galpha(16/z) chimeras efficiently link a wide range of G pro- Natl. Acad. Sci. USA 102, 892–897 (2005).
tein-coupled receptors to calcium mobilization. J. Biomol. Screen. 8, 39–49 123. Kenakin, T. Predicting therapeutic value in the lead optimization phase of drug
(2003). discovery. Nat. Rev. Drug Discov. 2, 429–438 (2003).
92. Kostenis, E. Is Galpha16 the optimal tool for fishing ligands of orphan G-protein- 124. Van Der Hee, R.M., Deurholt, T., Gerhardt, C.C. & De Groene, E.M. Comparison
coupled receptors? Trends Pharmacol. Sci. 22, 560–564 (2001). of 3 AT1 receptor binding assays: filtration assay, ScreenReady Target, and WGA
93. Trinquet, E. et al. D-myo-inositol 1-phosphate as a surrogate of D-myo-inositol Flashplate. J. Biomol. Screen. 10, 118–126 (2005).
1,4,5-tris phosphate to monitor G protein-coupled receptor activation. Anal. 125. Wu, X., Glickman, J.F., Bowen, B.R. & Sills, M.A. Comparison of assay technolo-
Biochem. 358, 126–135 (2006). gies for a nuclear receptor assay screen reveals differences in the sets of identi-
94. Hamdan, F.F., Audet, M., Garneau, P., Pelletier, J. & Bouvier, M. High-throughput fied functional antagonists. J. Biomol. Screen. 8, 381–392 (2003).
screening of G protein-coupled receptor antagonists using a bioluminescence 126. Inglese, J. et al. Quantitative high-throughput screening: a titration-based
resonance energy transfer 1-based beta-arrestin2 recruitment assay. J. Biomol. approach that efficiently identifies biological activities in large chemical librar-
Screen. 10, 463–475 (2005). ies. Proc. Natl. Acad. Sci. USA 103, 11473–11478 (2006).
95. Zheng, W., Spencer, R.H. & Kiss, L. High throughput assay technologies for ion 127. Burlingham, B. & Widlanski, T. An intuitive look at the relationship of Ki and
channel drug discovery. Assay Drug Dev. Technol. 2, 543–552 (2004). IC50: a more general use for the Dixon plot. J. Chem. Educ. 80, 214–218
96. McDonough, S.I. & Bean, B.P. in Voltage-Gated Ion Channels as Drug Targets (2003).
Vol. 29 (eds. Triggle, D.J., Gopalakrishnan, M., Rampe, D. & Zheng, W.) 19–35 128. Copeland, R.A. Mechanistic considerations in high-throughput screening. Anal.
(Wiley-VCH, Weinheim, Germany, 2006). Biochem. 320, 1–12 (2003).
97. Gee, K.R. et al. Chemical and physiological characterization of fluo-4 Ca(2+)- 129. Leff, P. & Dougall, I.G. Further concerns over Cheng-Prusoff analysis. Trends
indicator dyes. Cell Calcium 27, 97–106 (2000). Pharmacol. Sci. 14, 110–112 (1993).
98. Gonzalez, J.E. & Maher, M.P. Cellular fluorescent indicators and voltage/ion 130. Wu, G., Yuan, Y. & Hodge, C.N. Determining appropriate substrate conversion for
probe reader (VIPR) tools for ion channel and receptor drug discovery. Receptors enzymatic assays in high-throughput screening. J. Biomol. Screen. 8, 694–700
Channels 8, 283–295 (2002). (2003).
99. Terstappen, G.C. Nonradioactive rubidium ion efflux assay and its applications 131. Knight, S.M., Umezawa, N., Lee, H.S., Gellman, S.H. & Kay, B.K. A fluorescence
in drug discovery and development. Assay Drug Dev. Technol. 2, 553–559 polarization assay for the identification of inhibitors of the p53–DM2 protein-
(2004). protein interaction. Anal. Biochem. 300, 230–236 (2002).
100. Finkel, A. et al. Population patch clamp improves data consistency and success 132. Kane, S.A. et al. Development of a binding assay for p53/HDM2 by using homo-
rates in the measurement of ionic currents. J. Biomol. Screen. 11, 488–496 geneous time-resolved fluorescence. Anal. Biochem. 278, 29–38 (2000).
(2006). 133. Zuck, P. et al. Ligand-receptor binding measured by laser-scanning imaging. Proc.
101. Wible, B.A. et al. HERG-Lite: a novel comprehensive high-throughput screen for Natl. Acad. Sci. USA 96, 11122–11127 (1999).
drug-induced hERG risk. J. Pharmacol. Toxicol. Methods 52, 136–145 (2005). 134. Zhang, R. et al. Fluorescence polarization assay and inhibitor design for MDM2/
102. Schwab, D., Fischer, H., Tabatabaei, A., Poli, S. & Huwyler, J. Comparison of p53 interaction. Anal. Biochem. 331, 138–146 (2004).
in vitro P-glycoprotein screening assays: recommendations for their use in drug 135. Parks, D.J. et al. 1,4-Benzodiazepine-2,5-diones as small molecule antagonists
discovery. J. Med. Chem. 46, 1716–1725 (2003). of the HDM2-p53 interaction: discovery and SAR. Bioorg. Med. Chem. Lett. 15,
103. Henrich, C.J. et al. A high-throughput cell-based assay for inhibitors of ABCG2 765–770 (2005).
activity. J. Biomol. Screen. 11, 176–183 (2006). 136. Inglese, J. Engaging science and technology for drug development. Assay Drug
104. Wieczorke, R., Dlugai, S., Krampe, S. & Boles, E. Characterisation of mamma- Dev. Technol. 1, 1 (2002).
lian GLUT glucose transporters in a heterologous yeast expression system. Cell. 137. Lundholt, B.K., Heydorn, A., Bjorn, S.P. & Praestegaard, M. A simple cell-based
Physiol. Biochem. 13, 123–134 (2003). HTS assay system to screen for inhibitors of p53-Hdm2 protein-protein interac-
105. Verkman, A.S. Drug discovery in academia. Am. J. Physiol. Cell Physiol. 286, tions. Assay Drug Dev. Technol. 4, 679–688 (2006).
C465–C474 (2004). 138. Peppard, J. et al. Development of a high-throughput screening assay for inhibi-
106. Rubin, L.L. & de Sauvage, F.J. Targeting the Hedgehog pathway in cancer. Nat. tors of aggrecan cleavage using luminescent oxygen channeling (AlphaScreen).
Rev. Drug Discov. 5, 1026–1033 (2006). J. Biomol. Screen. 8, 149–156 (2003).
107. Massoud, T.F., Paulmurugan, R., De, A., Ray, P. & Gambhir, S.S. Reporter gene 139. DeForge, L.E. et al. Substrate capacity considerations in developing kinase
imaging of protein-protein interactions in living subjects. Curr. Opin. Biotechnol. assays. Assay Drug Dev. Technol. 2, 131–140 (2004).
18, 31–37 (2007). 140. Iyer, R., Barrese, A.A. III, Parakh, S., Parker, C.N. & Tripp, B.C. Inhibition profil-
108. MacDonald, M.L. et al. Identifying off-target effects and hidden phenotypes of ing of human carbonic anhydrase II by high-throughput screening of structur-
drugs in human cells. Nat. Chem. Biol. 2, 329–337 (2006). ally diverse, biologically active compounds. J. Biomol. Screen. 11, 782–791
109. Buttner, F.H., Kumpf, R., Menzel, S., Reulle, D. & Valler, M.J. Evaluation of the (2006).
InteraX system technology in a high-throughput screening environment. J. Biomol. 141. Comley, J. Assay Interference, a limiting factor in HTS? Drug Discov. World 4,
Screen. 10, 485–494 (2005). 91–98 (2003).
110. Wallrabe, H. & Periasamy, A. Imaging protein molecules using FRET and FLIM 142. Zheng, W. et al. Miniaturization of a hepatitis C virus RNA polymerase assay using
microscopy. Curr. Opin. Biotechnol. 16, 19–27 (2005). a -102 degrees C cooled CCD camera-based imaging system. Anal. Biochem. 290,
111. Marine, S. et al. A miniaturized cell-based fluorescence resonance energy transfer 214–220 (2001).
assay for insulin-receptor activation. Anal. Biochem. 355, 267–277 (2006). 143. Rishton, G.M. Nonleadlikeness and leadlikeness in biochemical screening. Drug
112. Wolff, M. et al. Automated high content screening for phosphoinositide 3 kinase Discov. Today 8, 86–96 (2003).
inhibition using an AKT 1 redistribution assay. Comb. Chem. High Throughput 144. Swift, K., Anderson, S.N. & Matayoshi, E.D. in Advances in Fluorescence Sensing
Screen. 9, 339–350 (2006). Technology V, Proceedings of SPIE - The International Society for Optical

478 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY


REVIEW

Engineering 4252 47–58 (2001). of particle binding kinetics by chemiluminescence. Proc. Natl. Acad. Sci. USA
145. Porcari, V., Magnoni, L., Terstappen, G.C. & Fecke, W. A continuous time- 91, 5426–5430 (1994).
resolved fluorescence assay for identification of BACE1 inhibitors. Assay Drug 150. Issad, T., Blanquart, C. & Gonzalez-Yanes, C. The use of bioluminescence reso-
Dev. Technol. 3, 287–297 (2005). nance energy transfer for the study of therapeutic targets: application to tyrosine
146. Moger, J., Gribbon, P., Sewing, A. & Winlove, C.P. The application of fluorescence kinase receptors. Expert Opin. Ther. Targets 11, 541–556 (2007).
lifetime readouts in high-throughput screening. J. Biomol. Screen. 11, 765–772 151. Zlokarnik, G. et al. Quantitation of transcription and clonal selection of single
(2006). living cells with beta-lactamase as reporter. Science 279, 84–88 (1998).
147. Feng, B.Y. et al. A high-throughput screen for aggregation-based inhibition in a 152. Kumaraswamy, S. et al. Fluorescent-conjugated polymer superquenching facili-
large compound library. J. Med. Chem. 50, 2385–2390 (2007). tates highly sensitive detection of proteases. Proc. Natl. Acad. Sci. USA 101,
148. Bakhtiarova, A. et al. Resveratrol inhibits firefly luciferase. Biochem. Biophys. 7511–7515 (2004).
Res. Commun. 351, 481–484 (2006). 153. Kiss, L. et al. High throughput ion-channel pharmacology: planar-array-based
149. Ullman, E.F. et al. Luminescent oxygen channeling immunoassay: measurement voltage clamp. Assay Drug Dev. Technol. 1, 127–135 (2003).

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 479


Supplementary Table 1: Assay Technologies Additional Entries
Refs. Refs.
Technology Description
Technology Assays
AAS Atomic absorption spectroscopy for measurements of ion transport/flux (primarily 1, 2 3-7
focused on K+ channels with Rb+ as tracer); 96/384-well microtiter plates
enabled with, for example, the ICR 12000.

Absorbance Follows Beer's Law; trans or epi-absorbance; via fluorescence attenuation 8, 9 10

ALPHA Amplified luminescent proximity homogeneous assay; donor-acceptor beads of 11, 12 13-16
~ 200 nm dia.;λex 680 nm (laser); singlet oxygen reacts with acceptor bead within
~200 nm distance; λem 520-620 nm

Bioluminescence Renilla or firefly luciferases used in reporter gene assays; analogs of D-luciferin 17 18-27
used to expand biochemical assays for proteases, P450s, monoamine oxidases
28, 29 30-37
Bla/CCF4 FRET-based assay using β-lactamase and β-lactam-containing substrate (e.g.
CCF4); λex 395 nm, λem 460/530 nm; ratiometric reporter gene assays and other
formats.

BRET Bioluminescence resonance energy transfer; useful for protein-protein 38, 39 40


interactions; Renilla luciferase donor luminescence (480 nm) stimulates
fluorescence of GFP/YFP acceptor; follows dipole-dipole interaction 1/r6 (~ 5 nm
maximum)

DELFIA Dissociation-enhanced lanthanide fluorescent immunoassay/Time resolved 41, 42 43-49


fluorescence; ELISA-like format using lanthanide chelate-labeled antibody; τf ~
500 usec; separation–based enhanced fluorescence; variations employing
labeled ligands; extremely sensitive.

ECL Electrochemiluminescence assay. Uses micro-carbon electrode plates allowing 50 20


electrical oxidation of a Ru(phen)32+ label. Multiplex applications in ELISA
format.
51, 52 53-58
EFC Enzyme fragment complementation. Based on β-galactosidase α-
complementation; High affinity complementation. Applied to both cell-free and
cell-based systems.

FCS / FIDA Fluorescence correlation spectroscopy / Fluorescence intensity distribution 59, 60 61-65
analysis; single molecule detection technology. Stochastic sampling of individual
labeled ligand fluorescence using confocal optics with illumination and detection
volume of ~1 femtoliter (fL), from a 1 uL sample volume. Enabled with the
Evotec Mark-II system.

FLT Fluorescent lifetime; time-domain method (~ns) measures the rate of fluorescent 66
decay following excitation; promises to eliminate fluorescence compound
interference; only recently applied to some microplate systems.
67, 68 69-78
FP Fluorescence polarization; polarized light reports change in probe ligand τc;
fluorophores of τf ~ 5 ns needed to measure ligand of MW < 1500 binding to a
receptor of MW ≥ 10,000; widely applied, ratiometric measurement.

FRET Fluorescent resonance energy transfer; donor-acceptor pair as in EDANS- 79, 80 81-84
DABCYL peptides or fluorescent proteins (e.g. CFP/YFP). Signal distance
dependence ~ 5 nm maximum.
(see also, Bla/CCF4)

Fluorescence Measurement of fluorescence intensity including the use of chromophores for 80, 85 86-97
quenching as in profluorescent protesase substrates (~ 5 nm maximum
distance); widely applied; superquenching varies in that non-equimolar amounts
of donor fluorophore and quencher are used. Includes kinetic modes of detection
with timescales ~ minutes to hours. Labeled antibodies extensively used in
imaging microscopy (see also ‘GFP and variants’ entry for references)

Fluorescence Requires fluorescent imaging plate reader or equivalent (e.g. FLIPR, FDS6000, 98, 99 99-108
VIPR); timescale ~ seconds; GPCR functional assays using Ca2+ sensitive dyes,
(rapid kinetic)
or steady-state ion channel activity using voltage sensitive dyes.
GFP and variants Cell-based imaging microscopy/cytometry assays; translocation, redistribution, 109-112 18, 110,
protein-protein interactions, reporter gene assays; HTS enabled with microplate-
113-118
compatible microscope-based imagers (e.q., ArrayScan, Opera, Pathway
BioImager, and INCell imagers) and cytometers (below).

HT High-throughput electrophysiology; measurement of ion channel currents with 119-122 123-127


microaperture arrays such as the planar patch-clamp electrode (e.g.,
Electrophysiology
PatchXpress 7000A, IonWorks Quattro). Transporter assays using solid
supported membranes (e.g., SURFE2R) (see also ion flux measurements via
AAS)
“Label Free” See following three entries
Spectrometry Mass spectrometry (e.g., RapidFire and SpeedScreen), high-field NMR (SAR by 128-131 132-135
NMR), fragment based methods using X-ray crystallography.

Optical Refractive index based. Measurement of ligand-binding using prism-coupled 136-138 139, 140
evanescent waves sensors, including surface plasmon resonance (SPR),
resonant waveguide grating (RWG), and resonant mirrors. kon and koff measured.
(e.g., Biacore, Plasmon Imager, IBIS, IAsys, Epic, Octet System). 384-well
microplate-based assays enabled with Epic system.

Non-optical Includes measurements of electrical impedance, quartz crystal acoustic 141-146 147-151
resonance, heat of binding /unfolding, ∆H°, and use of microcantilevers. Systems
for measuring electrical impedance (e.g., RT-CES) are particularly useful for cell-
based assays, and piezoelectric quartz crystal-based resonant acoustic profiling
(e.g., RAP♦id 4), isothermal (e.g. AutoITC) and differential scanning calorimetry
(e.g., VP-Capillary DSC System, MiDiCal) have generally been applied to
measure molecular interactions and enzyme reactions.

Laser Cytometry See following three entries


Acumen-enabled Commonly referred to as ‘Laser-scanning fluorescence microplate cytometry’ 152 153
Based on optical thresholding of fluorescence signal from cellular fluorescence
compared to background fluorescence. Used with adherent or settled cells,
suitable for microplate-based assays. Enabled with the Acumen Explorer eX3
(405, 488, and 633 nm laser excitation).
FMAT-enabled Fluorometric microvolume assay technology; also known as, Laser-scanning 154, 155 43, 156-162
imaging or laser-scanning cytometry. Relies on ‘macro-confocal’ optical
discrimination of fluorescence signal from accumulation of fluorescent ligands on
a receptor-containing cell or particle over that of unbound ligand in an equivalent
area of surrounding buffer. Used with adherent cells or settled beads, suitable
for microplate-based assays. Enabled with the 8200 Cellular Detection System,
formerly known as the FMAT (633 nm HeNe laser excitation).

Flow-based Flow cytometry detection using laser-induced fluorescence or light scatter from 163-165 166-168
individual cells or beads traveling sequentially in a hydrodynamically focused
laminar flow. The optical configuration permits limited excitation of sample buffer
surrounding the cell only, thus reducing the background signal from, for example,
unbound fluorescent ligand. Dye-encoded beads allow high level of multiplexing.

PCA Protein complementation assay; recombinant split reporters (e.g., GFP, 169, 170 153, 171,
luciferase, β-lactamase) fused to interacting proteins; phenotypic assays; low 172
affinity complementation.
SPA, FlashPlate Scintillation proximity assay; non-separation based radiometric detection using 173 174-186
solid supports; typical isotopes include 35S, 33P or 3H

ThermoFluor Temperature-dependent protein unfolding is monitored using environmentally 187 188


sensitive fluorescent probes that interact specifically with nonnative or unfolded
exposed hydrophobic regions of the target protein. Compounds that bind and
stabilize the protein will shift melting temperature, Tm, to a higher temperature,
destabilizing substances lower the Tm.

TR-FRET Time resolved-FRET; also known commercially as HTRF/LANCE; non- 41, 42, 79 44, 189-201
separation based format employing a lanthanide cryptate donor fluorophore of
long fluorescence lifetime ( τf ~ 500 usec); detections distance ~ 9 nm max.

NOTES: τf = fluorescent lifetime.τc = rotational correlation time.


References

1. Terstappen, G.C. Nonradioactive rubidium ion efflux assay and its applications in drug discovery and development. Assay Drug Dev
Technol 2, 553-559 (2004).
2. Stankovich, L., Wicks, D., Despotovski, S. & Liang, D. Atomic absorption spectroscopy in ion channel screening. Assay Drug Dev Technol
2, 569-574 (2004).
3. Parihar, A.S. et al. Functional analysis of large conductance Ca2(+)-activated K(+) channels: ion flux studies by atomic absorption
spectrometry. Assay Drug Dev Technol 1, 647-654 (2003).
4. Gill, S., Gill, R., Wicks, D., Despotovski, S. & Liang, D. Development of an HTS assay for Na+, K+-ATPase using nonradioactive rubidium
ion uptake. Assay Drug Dev Technol 2, 535-542 (2004).
5. Wang, K. et al. Validation of an atomic absorption rubidium ion efflux assay for KCNQ/M-channels using the ion Channel Reader 8000.
Assay Drug Dev Technol 2, 525-534 (2004).
6. Chaudhary, K.W. et al. Evaluation of the rubidium efflux assay for preclinical identification of HERG blockade. Assay Drug Dev Technol 4,
73-82 (2006).
7. Jow, F. et al. Rb+ efflux through functional activation of cardiac KCNQ1/minK channels by the benzodiazepine R-L3 (L-364,373). Assay
Drug Dev Technol 4, 443-450 (2006).
8. Lavery, P., Brown, M.J. & Pope, A.J. Simple absorbance-based assays for ultra-high throughput screening. J Biomol Screen 6, 3-9 (2001).
9. Zuck, P. et al. Miniaturization of absorbance assays using the fluorescent properties of white microplates. Anal Biochem 342, 254-259
(2005).
10. John, S., Fletcher, T.M., 3rd & Jonsson, C.B. Development and application of a high-throughput screening assay for HIV-1 integrase
enzyme activities. J Biomol Screen 10, 606-614 (2005).
11. Osmond, R.I. et al. GPCR screening via ERK 1/2: a novel platform for screening G protein-coupled receptors. J Biomol Screen 10, 730-
737 (2005).
12. Ullman, E.F. et al. Luminescent oxygen channeling immunoassay: measurement of particle binding kinetics by chemiluminescence. Proc
Natl Acad Sci U S A 91, 5426-5430 (1994).
13. Glickman, J.F. et al. A comparison of ALPHAScreen, TR-FRET, and TRF as assay methods for FXR nuclear receptors. Journal of
Biomolecular Screening 7, 3-10 (2002).
14. Peppard, J. et al. Development of a high-throughput screening assay for inhibitors of aggrecan cleavage using luminescent oxygen
channeling (AlphaScreen). J Biomol Screen 8, 149-156 (2003).
15. Wu, X., Glickman, J.F., Bowen, B.R. & Sills, M.A. Comparison of assay technologies for a nuclear receptor assay screen reveals
differences in the sets of identified functional antagonists. J Biomol Screen 8, 381-392 (2003).
16. Wu, X., Sills, M.A. & Zhang, J.H. Further comparison of primary hit identification by different assay technologies and effects of assay
measurement variability. J Biomol Screen 10, 581-589 (2005).
17. Fan, F. & Wood, K.V. Bioluminescent assays for high-throughput screening. Assay Drug Dev Technol 5, 127-136 (2007).
18. Bandyopadhyay, S. et al. A high-throughput drug screen targeted to the 5'untranslated region of Alzheimer amyloid precursor protein
mRNA. J Biomol Screen 11, 469-480 (2006).
19. Hao, W. et al. Development of a novel dicistronic reporter-selectable hepatitis C virus replicon suitable for high-throughput inhibitor
screening. Antimicrob Agents Chemother 51, 95-102 (2007).
20. Kenten, J.H. et al. Assays for high-throughput screening of e2 and e3 ubiquitin ligases. Methods Enzymol 399, 682-701 (2005).
21. O'Boyle, D.R., 2nd et al. Development of a cell-based high-throughput specificity screen using a hepatitis C virus-bovine viral diarrhea
virus dual replicon assay. Antimicrob Agents Chemother 49, 1346-1353 (2005).
22. Yen, L., Magnier, M., Weissleder, R., Stockwell, B.R. & Mulligan, R.C. Identification of inhibitors of ribozyme self-cleavage in mammalian
cells via high-throughput screening of chemical libraries. Rna 12, 797-806 (2006).
23. Branchini, B.R. et al. Thermostable red and green light-producing firefly luciferase mutants for bioluminescent reporter applications. Anal
Biochem 361, 253-262 (2007).
24. Jones, L.R. et al. Releasable luciferin-transporter conjugates: tools for the real-time analysis of cellular uptake and release. J Am Chem
Soc 128, 6526-6527 (2006).
25. Dupriez, V.J., Maes, K., Le Poul, E., Burgeon, E. & Detheux, M. Aequorin-based functional assays for G-protein-coupled receptors, ion
channels and tyrosine kinase receptors. Receptors and Channels 8, 319-330 (2002).
26. Wunder, F. et al. A cell-based cGMP assay useful for ultra-high-throughput screening and identification of modulators of the nitric
oxide/cGMP pathway. Anal Biochem 339, 104-112 (2005).
27. Jones, J. et al. High throughput drug screening for human immunodeficiency virus type 1 reactivating compounds. Assay Drug Dev
Technol 5, 181-189 (2007).
28. Zlokarnik, G. Fusions to beta-lactamase as a reporter for gene expression in live mammalian cells. Methods Enzymol 326, 221-244
(2000).
29. Zlokarnik, G. et al. Quantitation of transcription and clonal selection of single living cells with beta-lactamase as reporter. Science 279, 84-
88 (1998).
30. Stack, J.H., Whitney, M., Cubitt, A.B. & Pollok, B.A. in PCT Int. Appl. 171 pp. ((Aurora Biosciences Corporation, USA). Wo; 2001).
31. Oosterom, J., van Doornmalen, E.J., Lobregt, S., Blomenrohr, M. & Zaman, G.J. High-throughput screening using beta-lactamase
reporter-gene technology for identification of low-molecular-weight antagonists of the human gonadotropin releasing hormone receptor.
Assay Drug Dev Technol 3, 143-154 (2005).
32. Chin, J. et al. Miniaturization of cell-based beta-lactamase-dependent FRET assays to ultra-high throughput formats to identify agonists of
human liver X receptors. Assay Drug Dev Technol 1, 777-787 (2003).
33. Hobson, J.P., Liu, S., Rono, B., Leppla, S.H. & Bugge, T.H. Imaging specific cell-surface proteolytic activity in single living cells. Nat
Methods 3, 259-261 (2006).
34. Kunapuli, P. et al. Development of an intact cell reporter gene beta-lactamase assay for G protein-coupled receptors for high-throughput
screening. Anal Biochem 314, 16-29 (2003).
35. Peekhaus, N.T. et al. A beta-lactamase-dependent Gal4-estrogen receptor beta transactivation assay for the ultra-high throughput
screening of estrogen receptor beta agonists in a 3456-well format. Assay Drug Dev Technol 1, 789-800 (2003).
36. Whitney, M. et al. A collaborative screening program for the discovery of inhibitors of HCV NS2/3 cis-cleaving protease activity. J Biomol
Screen 7, 149-154 (2002).
37. Zuck, P. et al. A cell-based beta-lactamase reporter gene assay for the identification of inhibitors of hepatitis C virus replication. Anal
Biochem 334, 344-355 (2004).
38. Issad, T., Blanquart, C. & Gonzalez-Yanes, C. The use of bioluminescence resonance energy transfer for the study of therapeutic targets:
Application to tyrosine kinase receptors. Expert Opinion on Therapeutic Targets 11, 541-556 (2007).
39. Pfleger, K.D. & Eidne, K.A. Illuminating insights into protein-protein interactions using bioluminescence resonance energy transfer (BRET).
Nat Methods 3, 165-174 (2006).
40. Hamdan, F.F., Audet, M., Garneau, P., Pelletier, J. & Bouvier, M. High-throughput screening of G protein-coupled receptor antagonists
using a bioluminescence resonance energy transfer 1-based beta-arrestin2 recruitment assay. J Biomol Screen 10, 463-475 (2005).
41. Hemmila, I. & Laitala, V. Progress in lanthanides as luminescent probes. J Fluoresc 15, 529-542 (2005).
42. Trinquet, E. & Mathis, G. Fluorescence technologies for the investigation of chemical libraries. Mol Biosyst 2, 380-387 (2006).
43. Beasley, J.R., McCoy, P.M., Walker, T.L. & Dunn, D.A. Miniaturized, ultra-high throughput screening of tyrosine kinases using
homogeneous, competitive fluorescence immunoassays. Assay Drug Dev Technol 2, 141-151 (2004).
44. Gabriel, D. et al. High throughput screening technologies for direct cyclic AMP measurement. Assay Drug Dev Technol 1, 291-303 (2003).
45. Jin, G. et al. Development and comparison of nonradioactive in vitro kinase assays for NIMA-related kinase 2. Anal Biochem 358, 59-69
(2006).
46. Sadler, T.M. et al. Development and comparison of two nonradioactive kinase assays for I kappa B kinase. Anal Biochem 326, 106-113
(2004).
47. Waddleton, D., Ramachandran, C. & Wang, Q. Development of a time-resolved fluorescent assay for measuring tyrosine-phosphorylated
proteins in cells. Anal Biochem 309, 150-157 (2002).
48. Su, J.L. et al. A cell-based time-resolved fluorescence assay for selection of antibody reagents for G protein-coupled receptor
immunohistochemistry. J Immunol Methods 291, 123-135 (2004).
49. Inglese, J. et al. Chemokine receptor-ligand interactions measured using time-resolved fluorescence. Biochemistry 37, 2372-2377 (1998).
50. Ross, S. et al. High-content screening analysis of the p38 pathway: profiling of structurally related p38alpha kinase inhibitors using cell-
based assays. Assay Drug Dev Technol 4, 397-409 (2006).
51. Eglen, R.M. & Singh, R. Beta galactosidase enzyme fragment complementation as a novel technology for high throughput screening.
Comb Chem High Throughput Screen 6, 381-387 (2003).
52. Wehrman, T.S., Casipit, C.L., Gewertz, N.M. & Blau, H.M. Enzymatic detection of protein translocation. Nat Methods 2, 521-527 (2005).
53. Golla, R. & Seethala, R. A homogeneous enzyme fragment complementation cyclic AMP screen for GPCR agonists. Journal of
Biomolecular Screening 7, 515-525 (2002).
54. Zhao, X. et al. Homogeneous assays for cellular protein degradation using beta-galactosidase complementation: NF-kappaB/IkappaB
pathway signaling. Assay Drug Dev Technol 1, 823-833 (2003).
55. Weber, M. et al. A 1536-well cAMP assay for Gs- and Gi-coupled receptors using enzyme fragmentation complementation. Assay Drug
Dev Technol 2, 39-49 (2004).
56. Naqvi, T., Lim, A., Rouhani, R., Singh, R. & Eglen, R.M. Beta galactosidase enzyme fragment complementation as a high-throughput
screening protease technology. J Biomol Screen 9, 398-408 (2004).
57. Weber, M. et al. Ultra-High-Throughput Screening for Antagonists of A Gi-Coupled Receptor in A 2.2-mul 3,456-Well Plate Format
CyclicAMP Assay. Assay Drug Dev Technol 5, 117-126 (2007).
58. Zaman, G.J., van der Lee, M.M., Kok, J.J., Nelissen, R.L. & Loomans, E.E. Enzyme fragment complementation binding assay for
p38alpha mitogen-activated protein kinase to study the binding kinetics of enzyme inhibitors. Assay Drug Dev Technol 4, 411-420 (2006).
59. Heilker, R., Zemanova, L., Valler, M.J. & Nienhaus, G.U. Confocal fluorescence microscopy for high-throughput screening of G-protein
coupled receptors. Curr Med Chem 12, 2551-2559 (2005).
60. Gribbon, P. et al. Experiences in implementing uHTS--cutting edge technology meets the real world. Curr Drug Discov Technol 1, 27-35
(2004).
61. Wolcke, J., Hunt, N., Jungmann, J. & Ullmann, D. Early identification of false positives in high-throughput screening for activators of p53-
DNA interaction. J Biomol Screen 11, 341-350 (2006).
62. Moore, K.J. et al. Single Molecule Detection Technologies in Miniaturized High Throughput Screening: Fluorescence Correlation
Spectroscopy. J Biomol Screen 4, 335-354 (1999).
63. Koltermann, A., Kettling, U., Bieschke, J., Winkler, T. & Eigen, M. Rapid assay processing by integration of dual-color fluorescence cross-
correlation spectroscopy: high throughput screening for enzyme activity. Proc Natl Acad Sci U S A 95, 1421-1426 (1998).
64. Rudiger, M., Haupts, U., Moore, K.J. & Pope, A.J. Single-molecule detection technologies in miniaturized high throughput screening:
binding assays for G protein-coupled receptors using fluorescence intensity distribution analysis and fluorescence anisotropy. Journal of
Biomolecular Screening 6, 29-37 (2001).
65. Scheel, A.A. et al. Receptor-ligand interactions studied with homogeneous fluorescence-based assays suitable for miniaturized screening.
Journal of Biomolecular Screening 6, 11-18 (2001).
66. Moger, J., Gribbon, P., Sewing, A. & Winlove, C.P. The application of fluorescence lifetime readouts in high-throughput screening. J
Biomol Screen 11, 765-772 (2006).
67. Combinatorial Chemistry & High Throughput Screening 6 (2003).
68. Owicki, J.C. Fluorescence polarization and anisotropy in high throughput screening: perspectives and primer. J Biomol Screen 5, 297-306
(2000).
69. Turek-Etienne, T.C. et al. Evaluation of fluorescent compound interference in 4 fluorescence polarization assays: 2 kinases, 1 protease,
and 1 phosphatase. Journal of Biomolecular Screening 8, 176-184 (2003).
70. Banks, P. & Harvey, M. Considerations for using fluorescence polarization in the screening of G protein-coupled receptors. Journal of
Biomolecular Screening 7, 111-117 (2002).
71. Sportsman, J.R., Gaudet, E.A. & Boge, A. Immobilized metal ion affinity-based fluorescence polarization (IMAP): advances in kinase
screening. Assay Drug Dev Technol 2, 205-214 (2004).
72. Vedvik, K.L. et al. Overcoming compound interference in fluorescence polarization-based kinase assays using far-red tracers. Assay Drug
Dev Technol 2, 193-203 (2004).
73. Turek-Etienne, T.C., Kober, T.P., Stafford, J.M. & Bryant, R.W. Development of a fluorescence polarization AKT serine/threonine kinase
assay using an immobilized metal ion affinity-based technology. Assay Drug Dev Technol 1, 545-553 (2003).
74. Loomans, E.E., van Doornmalen, A.M., Wat, J.W. & Zaman, G.J. High-throughput screening with immobilized metal ion affinity-based
fluorescence polarization detection, a homogeneous assay for protein kinases. Assay Drug Dev Technol 1, 445-453 (2003).
75. Zhang, R. et al. Fluorescence polarization assay and inhibitor design for MDM2/p53 interaction. Anal Biochem 331, 138-146 (2004).
76. Parker, G.J., Law, T.L., Lenoch, F.J. & Bolger, R.E. Development of high throughput screening assays using fluorescence polarization:
nuclear receptor-ligand-binding and kinase/phosphatase assays. J Biomol Screen 5, 77-88 (2000).
77. Singh, P., Lillywhite, B., Bannaghan, C. & Broad, P. Using IMAP technology to identify kinase inhibitors: comparison with a substrate
depletion approach and analysis of the nature of false positives. Comb Chem High Throughput Screen 8, 319-325 (2005).
78. Lowery, R.G. & Kleman-Leyer, K. Transcreener: screening enzymes involved in covalent regulation. Expert Opin Ther Targets 10, 179-
190 (2006).
79. Comley, J. Kinase screening and profiling-spoilt for choice. Drug Discov World 7, 27-50 (2006).
80. Lakowicz, J.R. Principles of Fluorescence Spectroscopy. (Springer, Berlin; 2006).
81. Rodems, S.M. et al. A FRET-based assay platform for ultra-high density drug screening of protein kinases and phosphatases. Assay and
Drug Development Technologies 1, 9-19 (2002).
82. Hamman, B.D., Pollok, B.A., Bennett, T., Allen, J. & Heim, R. in Journal of Biomolecular Screening, Vol. 7 45-552002).
83. Jones, J., Heim, R., Hare, E., Stack, J. & Pollok, B.A. Development and application of a GFP-FRET intracellular caspase assay for drug
screening. Journal of Biomolecular Screening 5, 307-317 (2000).
84. Grant, S.K., Sklar, J.G. & Cummings, R.T. Development of novel assays for proteolytic enzymes using rhodamine-based fluorogenic
substrates. J Biomol Screen 7, 531-540 (2002).
85. Kumaraswamy, S. et al. Fluorescent-conjugated polymer superquenching facilitates highly sensitive detection of proteases. PNAS 101,
7511-7515 (2004).
86. Johnston, P.A. et al. Development and Implementation of a 384-well Homogeneous Fluorescence Intensity High-Throughput Screening
Assay to Identify Mitogen-Activated Protein Kinase Phosphatase-1 Dual-Specificity Phosphatase Inhibitors. Assay and Drug Development
Technologies 5, 319-332 (2007).
87. Buehler, C. et al. Multi-photon excitation of intrinsic protein fluorescence and its application to pharmaceutical drug screening. Assay Drug
Dev Technol 3, 155-167 (2005).
88. Billich, A. et al. Confocal fluorescence detection expanded to UV excitation: the first continuous fluorimetric assay of human steroid
sulfatase in nanoliter volume. Assay Drug Dev Technol 2, 21-30 (2004).
89. Baniecki, M.L., Wirth, D.F. & Clardy, J. High-throughput Plasmodium falciparum growth assay for malaria drug discovery. Antimicrob
Agents Chemother 51, 716-723 (2007).
90. Bhat, J. et al. High-throughput screening of RNA polymerase inhibitors using a fluorescent UTP analog. J Biomol Screen 11, 968-976
(2006).
91. Gribbon, P. & Sewing, A. Fluorescence readouts in HTS: no gain without pain? Drug Discov Today 8, 1035-1043 (2003).
92. Kashem, M.A. et al. Three mechanistically distinct kinase assays compared: Measurement of intrinsic ATPase activity identified the most
comprehensive set of ITK inhibitors. J Biomol Screen 12, 70-83 (2007).
93. Lim, M.H., Xu, D. & Lippard, S.J. Visualization of nitric oxide in living cells by a copper-based fluorescent probe. Nat Chem Biol 2, 375-380
(2006).
94. Peppard, J. et al. Development of an assay suitable for high-throughput screening to measure matrix metalloprotease activity. Assay Drug
Dev Technol 1, 425-433 (2003).
95. Tomazzolli, R., Serra, M.D., Bellisola, G., Colombatti, M. & Guella, G. A fluorescence-based assay for the reductase activity of protein
disulfide isomerase. Anal Biochem 350, 105-112 (2006).
96. Richards, G.R. et al. A morphology- and kinetics-based cascade for human neural cell high content screening. Assay Drug Dev Technol 4,
143-152 (2006).
97. Bertelsen, M. & Sanfridson, A. Inflammatory pathway analysis using a high content screening platform. Assay Drug Dev Technol 3, 261-
271 (2005).
98. Gonzalez, J.E. & Maher, M.P. Cellular fluorescent indicators and voltage/ion probe reader (VIPR) tools for ion channel and receptor drug
discovery. Receptors Channels 8, 283-295 (2002).
99. Chambers, C. et al. Measuring intracellular calcium fluxes in high throughput mode. Combinatorial Chemistry and High Throughput
Screening 6, 355-362 (2003).
100. Baxter, D.F. et al. A novel membrane potential-sensitive fluorescent dye improves cell-based assays for ion channels. J Biomol Screen 7,
79-85 (2002).
101. Niedernberg, A., Tunaru, S., Blaukat, A., Harris, B. & Kostenis, E. Comparative analysis of functional assays for characterization of agonist
ligands at G protein-coupled receptors. Journal of Biomolecular Screening 8, 500-510 (2003).
102. Hodder, P., Mull, R., Cassaday, J., Berry, K. & Strulovici, B. Miniaturization of intracellular calcium functional assays to 1536-well plate
format using a fluorometric imaging plate reader. J Biomol Screen 9, 417-426 (2004).
103. Hodder, P. et al. Identification of metabotropic glutamate receptor antagonists using an automated high-throughput screening system.
Anal Biochem 313, 246-254 (2003).
104. Wolff, C., Fuks, B. & Chatelain, P. Comparative study of membrane potential-sensitive fluorescent probes and their use in ion channel
screening assays. J Biomol Screen 8, 533-543 (2003).
105. Sullivan, E., Tucker, E.M. & Dale, I.L. Measurement of [Ca2+] using the Fluorometric Imaging Plate Reader (FLIPR). Methods Mol Biol
114, 125-133 (1999).
106. Lubin, M.L. et al. A nonadherent cell-based HTS assay for N-type calcium channel using calcium 3 dye. Assay Drug Dev Technol 4, 689-
694 (2006).
107. Felix, J.P. et al. Functional assay of voltage-gated sodium channels using membrane potential-sensitive dyes. Assay Drug Dev Technol 2,
260-268 (2004).
108. Hamelin, M. et al. A high-throughput assay for modulators of ligand-gated chloride channels. Assay Drug Dev Technol 3, 59-64 (2005).
109. Giepmans, B.N., Adams, S.R., Ellisman, M.H. & Tsien, R.Y. The fluorescent toolbox for assessing protein location and function. Science
312, 217-224 (2006).
110. Wolff, M. et al. Automated high content screening for phosphoinositide 3 kinase inhibition using an AKT 1 redistribution assay. Comb
Chem High Throughput Screen 9, 339-350 (2006).
111. Inglese, J. (ed.) Measuring Biological Responses with Automated Microscopy, Vol. 414. (Elsevier Acadmeic Press, San Diego; 2006).
112. Taylor, D.L., Haskins, J.R. & Giuliano, K.A. (eds.) High Content Screening, Vol. 356. (Humana Press, 2006).
113. Granas, C. et al. Identification of RAS-mitogen-activated protein kinase signaling pathway modulators in an ERF1 redistribution screen. J
Biomol Screen 11, 423-434 (2006).
114. Hudson, C.C., Oakley, R.H., Cruickshank, R.D., Rhem, S.M. & Loomis, C.R. Automation and validation of the Transfluor technology: a
universal screening assay for G protein-coupled receptors. Proceedings of SPIE-The International Society for Optical Engineering 4626,
548-555 (2002).
115. Lundholt, B.K., Heydorn, A., Bjorn, S.P. & Praestegaard, M. A simple cell-based HTS assay system to screen for inhibitors of p53-Hdm2
protein-protein interactions. Assay Drug Dev Technol 4, 679-688 (2006).
116. Oakley, R.H. et al. The cellular distribution of fluorescently labeled arrestins provides a robust, sensitive, and universal assay for screening
G protein-coupled receptors. Assay and Drug Development Technologies 1, 21-30 (2002).
117. Stubbs, S. & Thomas, N. Dynamic green fluorescent protein sensors for high-content analysis of the cell cycle. Methods Enzymol 414, 1-
21 (2006).
118. Zanella, F. et al. An HTS Approach to Screen for Antagonists of the Nuclear Export Machinery Using High Content Cell-Based Assays.
Assay and Drug Development Technologies 5, 333-341 (2007).
119. Schroeder, K., Neagle, B., Trezise, D.J. & Worley, J. Ionworks HT: a new high-throughput electrophysiology measurement platform. J
Biomol Screen 8, 50-64 (2003).
120. Kiss, L. et al. High throughput ion-channel pharmacology: planar-array-based voltage clamp. Assay Drug Dev Technol 1, 127-135 (2003).
121. Bennett, P.B. & Guthrie, H.R. Trends in ion channel drug discovery: advances in screening technologies. Trends Biotechnol 21, 563-569
(2003).
122. Kelety, B. et al. Transporter assays using solid supported membranes: a novel screening platform for drug discovery. Assay Drug Dev
Technol 4, 575-582 (2006).
123. Sorota, S., Zhang, X.S., Margulis, M., Tucker, K. & Priestley, T. Characterization of a hERG screen using the IonWorks HT: comparison to
a hERG rubidium efflux screen. Assay Drug Dev Technol 3, 47-57 (2005).
124. Tao, H. et al. Automated tight seal electrophysiology for assessing the potential hERG liability of pharmaceutical compounds. Assay Drug
Dev Technol 2, 497-506 (2004).
125. Dubin, A.E. et al. Identifying modulators of hERG channel activity using the PatchXpress planar patch clamp. J Biomol Screen 10, 168-
181 (2005).
126. Powell, A. in Satellite Conference in the 51th Annual Biophysics MeetingBaltimore, MD; 2007).
127. Finkel, A. et al. Population patch clamp improves data consistency and success rates in the measurement of ionic currents. J Biomol
Screen 11, 488-496 (2006).
128. Quercia, A.K. et al. High-throughput screening by mass spectrometry: comparison with the scintillation proximity assay with a focused-file
screen of AKT1/PKBalpha. J Biomol Screen 12, 473-480 (2007).
129. Stockman, B.J. & Dalvit, C. NMR screening techniques in drug discovery and drug design. Progress in Nuclear Magnetic Resonance
Spectroscopy 41, 187–231 (2002).
130. Hajduk, P.J. & Greer, J. A decade of fragment-based drug design: strategic advances and lessons learned. Nat Rev Drug Discov 6, 211-
219 (2007).
131. Pellecchia, M. et al. NMR-based techniques in the hit identification and optimisation processes. Expert Opin Ther Targets 8, 597-611
(2004).
132. Fejzo, J., Lepre, C. & Xie, X. Application of NMR screening in drug discovery. Curr Top Med Chem 3, 81-97 (2003).
133. Ozbal, C.C. et al. High throughput screening via mass spectrometry: a case study using acetylcholinesterase. Assay Drug Dev Technol 2,
373-381 (2004).
134. Zehender, H., Le Goff, F., Lehmann, N., Filipuzzi, I. & Mayr, L.M. SpeedScreen: The "missing link" between genomics and lead discovery.
J Biomol Screen 9, 498-505 (2004).
135. Vazquez, J. et al. Development of molecular probes for second-site screening and design of protein tyrosine phosphatase inhibitors. J
Med Chem 50, 2137-2143 (2007).
136. Fang, Y., Lahiri, J. & Picard, L. G protein-coupled receptor microarrays for drug discovery. Drug Discovery Today 8, 755-761 (2003).
137. Verdonk, E. et al. Cellular dielectric spectroscopy: a label-free comprehensive platform for functional evaluation of endogenous receptors.
Assay Drug Dev Technol 4, 609-619 (2006).
138. Rich, R.L. & Myszka, D.G. Higher-throughput, label-free, real-time molecular interaction analysis. Anal Biochem 361, 1-6 (2007).
139. Cunningham, B.T. et al. Label-free assays on the BIND system. J Biomol Screen 9, 481-490 (2004).
140. O'Malley, S.M., Xie, X. & Frutos, A.G. Label-free high-throughput functional lytic assays. J Biomol Screen 12, 117-125 (2007).
141. Hug, T.S. Biophysical methods for monitoring cell-substrate interactions in drug discovery. Assay Drug Dev Technol 1, 479-488 (2003).
142. Katz, E. & Willner, I. Probing Biomolecular Interactions at Conductive and Semiconductive Surfaces by Impedance Spectroscopy: Routes
to Impedimetric Immunosensors, DNA-Sensors, and Enzyme Biosensors. Electroanalysis 15, 913–947 (2003).
143. Li, X., Thompson, K.S., Godber, B. & Cooper, M.A. Quantification of small molecule-receptor affinities and kinetics by acoustic profiling.
Assay Drug Dev Technol 4, 565-573 (2006).
144. Atienza, J.M. et al. Dynamic and label-free cell-based assays using the real-time cell electronic sensing system. Assay Drug Dev Technol
4, 597-607 (2006).
145. Vermeir, S. et al. Microplate Differential Calorimetric Biosensor for Ascorbic Acid Analysis in Food and Pharmaceuticals. Anal Chem
(2007).
146. Plotnikov, V. et al. An autosampling differential scanning calorimeter instrument for studying molecular interactions. Assay Drug Dev
Technol 1, 83-90 (2002).
147. Yu, N. et al. Real-time monitoring of morphological changes in living cells by electronic cell sensor arrays: an approach to study G protein-
coupled receptors. Anal Chem 78, 35-43 (2006).
148. Kirstein, S.L. et al. Live cell quality control and utility of real-time cell electronic sensing for assay development. Assay Drug Dev Technol
4, 545-553 (2006).
149. Cooper, M.A. Non-optical screening platforms: the next wave in label-free screening? Drug Discov Today 11, 1068-1074 (2006).
150. Godber, B. et al. Direct quantification of analyte concentration by resonant acoustic profiling. Clin Chem 51, 1962-1972 (2005).
151. Wu, G. et al. Origin of nanomechanical cantilever motion generated from biomolecular interactions. Proc Natl Acad Sci U S A 98, 1560-
1564 (2001).
152. Bowen, W.P. & Wylie, P.G. Application of Laser-Scanning Fluorescence Microplate Cytometry in High Content Screening. Assay Drug
Dev Technol 4, 209-221 (2006).
153. Auld, D.S. et al. Fluorescent protein-based cellular assays analyzed by laser-scanning microplate cytometry in 1536-well plate format.
Methods Enzymol 414, 566-589 (2006).
154. Zuck, P. et al. Ligand-receptor binding measured by laser-scanning imaging. Proceedings of the National Academy of Sciences of the
United States of America 96, 11122-11127 (1999).
155. Miraglia, S. et al. Homogeneous Cell- and Bead-Based Assays for High Throughput Screening Using Fluorometric Microvolume Assay
Technology. J Biomol Screen 4, 193-204 (1999).
156. Auld, D.S., Dunn, D.A., Lehrach, J.M., McCoy, P. & Swanson, R. 1,536-well assay development and screening using whole cell
displacement binding and laser scanning imaging. Assay and Drug Development Technologies 1, 167-174 (2003).
157. Grand-Perret, T. et al. Assay development for high throughput screening of p21(Waf1/Cip1) protein expression in intact cells using
fluorometric microvolume assay technology. Drugs Exp Clin Res 30, 89-98 (2004).
158. Toney, J.H., Ogawa, A., Blair, M. & Park, Y.W. A "mix and read" assay for insulin using fluorometric microvolume assay technology. Assay
Drug Dev Technol 1, 521-525 (2003).
159. Lee, J.Y. et al. Oncology drug discovery applications using the FMAT 8100 HTS system. J Biomol Screen 8, 81-88 (2003).
160. Mellentin-Michelotti, J. et al. Determination of ligand binding affinities for endogenous seven-transmembrane receptors using fluorometric
microvolume assay technology. Anal Biochem 272, 182-190 (1999).
161. Swartzman, E.E., Miraglia, S.J., Mellentin-Michelotti, J., Evangelista, L. & Yuan, P.M. A homogeneous and multiplexed immunoassay for
high-throughput screening using fluorometric microvolume assay technology. Anal Biochem 271, 143-151 (1999).
162. Miller, D.K. et al. Development of a high-capacity homogeneous fluorescent assay for the measurement of leukotriene B4. Anal Biochem
349, 129-135 (2006).
163. Edwards, B.S., Oprea, T., Prossnitz, E.R. & Sklar, L.A. Flow cytometry for high-throughput, high-content screening. Curr Opin Chem Biol
8, 392-398 (2004).
164. Waller, A., Simons, P., Prossnitz, E.R., Edwards, B.S. & Sklar, L.A. High throughput screening of G-protein coupled receptors via flow
cytometry. Combinatorial Chemistry and High Throughput Screening 6, 389-397 (2003).
165. Iannone, M.A. et al. Multiplexed molecular interactions of nuclear receptors using fluorescent microspheres. Cytometry 44, 326-337
(2001).
166. Edwards, B.S. et al. Biomolecular screening of formylpeptide receptor ligands with a sensitive, quantitative, high-throughput flow cytometry
platform. Nat Protoc 1, 59-66 (2006).
167. Saunders, M.J. et al. Microsphere-based protease assays and screening application for lethal factor and factor Xa. Cytometry A 69, 342-
352 (2006).
168. Simons, P.C. et al. Duplexed, bead-based competitive assay for inhibitors of protein kinases. Cytometry A 71, 451-459 (2007).
169. MacDonald, M.L. et al. Identifying off-target effects and hidden phenotypes of drugs in human cells. Nat Chem Biol 2, 329-337 (2006).
170. Remy, I. & Michnick, S.W. Mapping biochemical networks with protein-fragment complementation assays. Methods Mol Biol 261, 411-426
(2004).
171. Yu, H. et al. Measuring drug action in the cellular context using protein-fragment complementation assays. Assay Drug Dev Technol 1,
811-822 (2003).
172. Michnick, S.W., MacDonald, M.L. & Westwick, J.K. Chemical genetic strategies to delineate MAP kinase signaling pathways using protein-
fragment complementation assays (PCA). Methods 40, 287-293 (2006).
173. Wu, S. & Liu, B. Application of scintillation proximity assay in drug discovery. BioDrugs 19, 383-392 (2005).
174. Ferrer, M. et al. A fully automated [35S]GTPgammaS scintillation proximity assay for the high-throughput screening of Gi-linked G protein-
coupled receptors. Assay Drug Dev Technol 1, 261-273 (2003).
175. Hui, X. et al. A robust homogeneous binding assay for alpha4beta2 nicotinic acetylcholine receptor. Acta Pharmacol Sin 26, 1175-1180
(2005).
176. Rodgers, G. et al. Development of displacement binding and GTPgammaS scintillation proximity assays for the identification of
antagonists of the micro-opioid receptor. Assay Drug Dev Technol 1, 627-636 (2003).
177. Zheng, W. et al. Miniaturization of a hepatitis C virus RNA polymerase assay using a -102 degrees C cooled CCD camera-based imaging
system. Anal Biochem 290, 214-220 (2001).
178. Crane, K. & Shih, D.T. Development of a homogeneous binding assay for histamine receptors. Anal Biochem 335, 42-49 (2004).
179. Liu, J.J., Hartman, D.S. & Bostwick, J.R. An immobilized metal ion affinity adsorption and scintillation proximity assay for receptor-
stimulated phosphoinositide hydrolysis. Anal Biochem 318, 91-99 (2003).
180. Glickman, J.F. & Schmid, A. Farnesyl pyrophosphate synthase: real-time kinetics and inhibition by nitrogen-containing bisphosphonates in
a scintillation assay. Assay Drug Dev Technol 5, 205-214 (2007).
181. Barrett, C. et al. Configuration of a scintillation proximity assay for the activity assessment of recombinant human adenine
phosphoribosyltransferase. Assay Drug Dev Technol 4, 661-669 (2006).
182. Bembenek, M.E. et al. A homogeneous scintillation proximity format for monitoring the activity of recombinant human long-chain-fatty-acyl-
CoA synthetase 5. Assay Drug Dev Technol 2, 300-307 (2004).
183. Bembenek, M.E. et al. Characterization of the kinase domain of the ephrin-B3 receptor tyrosine kinase using a scintillation proximity
assay. Assay Drug Dev Technol 1, 555-563 (2003).
184. Bembenek, M.E. et al. Development of a high-throughput assay for two inositol-specific phospholipase Cs using a scintillation proximity
format. Assay Drug Dev Technol 1, 435-443 (2003).
185. Weiss, D.R. & Glickman, J.F. Characterization of fatty acid synthase activity using scintillation proximity. Assay Drug Dev Technol 1, 161-
166 (2003).
186. Zhang, J. et al. A scintillation proximity assay for human interleukin-5 (hIL-5) high-affinity binding in insect cells coexpressing hIL-5
receptor alpha and beta subunits. Anal Biochem 268, 134-142 (1999).
187. Pantoliano, M.W. et al. High-density miniaturized thermal shift assays as a general strategy for drug discovery. J Biomol Screen 6, 429-
440 (2001).
188. Grasberger, B.L. et al. Discovery and cocrystal structure of benzodiazepinedione HDM2 antagonists that activate p53 in cells. J Med
Chem 48, 909-912 (2005).
189. Ferrer, M., Hamilton, A.C. & Inglese, J. A PDZ Domain-Based Detection System for Enzymatic Assays. Analytical Biochemistry 301, 207-
216 (2002).
190. Cummings, R.T., McGovern, H.M., Zheng, S., Park, Y.W. & Hermes, J.D. Use of a Phosphotyrosine-Antibody Pair as a General Detection
Method in Homogeneous Time-Resolved Fluorescence: Application to Human Immunodeficiency Viral Protease. Analytical Biochemistry
269, 79-93 (1999).
191. Zhou, G., Cummings, R., Hermes, J. & Moller, D.E. Use of Homogeneous Time-Resolved Fluorescence Energy Transfer in the
Measurement of Nuclear Receptor Activation. Methods (San Diego, CA, United States) 25, 54-61 (2001).
192. Ferrer, M. et al. Miniaturizable homogenous time-resolved fluorescence assay for carboxypeptidase B activity. Anal Biochem 317, 94-98
(2003).
193. Kennedy, M.E. et al. Measuring human beta-secretase (BACE1) activity using homogeneous time-resolved fluorescence. Anal Biochem
319, 49-55 (2003).
194. Leblanc, V. et al. Homogeneous time-resolved fluorescence assay for identifying p53 interactions with its protein partners, directly in a
cellular extract. Anal Biochem 308, 247-254 (2002).
195. Zhang, J.H., Chen, T., Nguyen, S.H. & Oldenburg, K.R. A high-throughput homogeneous assay for reverse transcriptase using generic
reagents and time-resolved fluorescence detection. Anal Biochem 281, 182-186 (2000).
196. Pope, A.J. Introduction LANCEtrade mark vs. HTRF(R) Technologies (or Vice Versa). J Biomol Screen 4, 301-302 (1999).
197. Trinquet, E. et al. D-myo-inositol 1-phosphate as a surrogate of D-myo-inositol 1,4,5-tris phosphate to monitor G protein-coupled receptor
activation. Anal Biochem 358, 126-135 (2006).
198. Preaudat, M. et al. A homogeneous caspase-3 activity assay using HTRF technology. J Biomol Screen 7, 267-274 (2002).
199. Ohmi, N., Wingfield, J.M., Yazawa, H. & Inagaki, O. Development of a homogeneous time-resolved fluorescence assay for high
throughput screening to identify Lck inhibitors: comparison with scintillation proximity assay and streptavidin-coated plate assay. J Biomol
Screen 5, 463-470 (2000).
200. Enomoto, K. et al. High throughput screening for human interferon-gamma production inhibitor using homogenous time-resolved
fluorescence. J Biomol Screen 5, 263-268 (2000).
201. Lundin, K., Blomberg, K., Nordstrom, T. & Lindqvist, C. Development of a time-resolved fluorescence resonance energy transfer assay
(cell TR-FRET) for protein detection on intact cells. Anal Biochem 299, 92-97 (2001).
Supplementary Table 2. Assays developed to measure the p53-DM2 interaction

Format Technology Components Conditions / Comment Ref:


p53 DM2
Competition FP FITC-[15-30]p53 peptidea GST-[9-149]DM2 • 15 nM FITC peptide 1
Binding (Kd, app =2 uM); low • 1 uM DM2
affinity peptide • 384-well plates
FITC-11, 11 amino acid • Non-Langmuir binding
phage peptide (Kd, app =1
uM)
Competition FP 6-FAM, 6 amino acid Trx-His6-factor Xa • 1 nM FAM peptide 2
Binding p53-derivedb peptide (Kd cleavage site-[17- • 10 nM DM2
= 2 nM) 125]DM2 • 384-well plates
• Non-Langmuir binding.
• Cubic equation to determine Kic
Competition TR-FRET biotin-trx-[1-83]p53; GST-[1-188]DM2; • 10 nM p53 3
Binding streptavidin (SA)-XL665 Eu(K)-anti-GST • 5 nM DM2
• 100 nM SA-XL665,
• 20K B-counts Eu(K)-anti-GST
• 96-well plates
• 35-fold rightward shift in unlabeled
[1-83]p53 IC50d
Phenotypic ‘GRIP’ Protein EGFP-[1-312]p53 PDE4A4-[1-124]DM2 • Internal system controls and Nutlin- 4
distributione 3 assay positive controls.f
• 384-well plates
Phenotypic PCA p53-Venus[2] Venus[1]-DM2 • Transient transfection 5
• 96-well plates

Direct Tm shift not applicable Glycine-[17-125]DM2 • 10 uM DM2 6


Binding ‘ThermoFluor’ • 100 uM fluorescent probe, Dapoxyl
sulfonic acid
• 384-well plates
• FP assay used to confirm actives
Notes:

a. Sequence of 15 amino acid p53 peptide = SQETFSDLWKLLPENN; phage display peptide= SFTDYWRDLEQ
b. Ac-FR-Dpr(6FAM)-Ac6c-(6-Br)WEEL-NH2
c. Uses the physiologically relevant cubic equation to determine Ki values in competition experiment under concentration depletion conditions of
the three starting species (e.g., non-Langmuir binding conditions).
d. Enhanced complex avidity, presumably multivalent interactions.
e. GRIP is an acronym for ‘Green fluorescent protein-assisted readout for interacting proteins.’
f. System controls: RS25344 induces localization of PDE4A4 into compact foci; RP73401 mediates foci dispersion.

References:

1. Knight, S.M., Umezawa, N., Lee, H.S., Gellman, S.H. & Kay, B.K. A fluorescence polarization assay for the identification of inhibitors of the
p53-DM2 protein-protein interaction. Anal Biochem 300, 230-236 (2002).
2. Zhang, R. et al. Fluorescence polarization assay and inhibitor design for MDM2/p53 interaction. Anal Biochem 331, 138-146 (2004).
3. Kane, S.A. et al. Development of a binding assay for p53/HDM2 by using homogeneous time-resolved fluorescence. Anal Biochem 278,
29-38 (2000).
4. Lundholt, B.K., Heydorn, A., Bjorn, S.P. & Praestegaard, M. A simple cell-based HTS assay system to screen for inhibitors of p53-Hdm2
protein-protein interactions. Assay Drug Dev Technol 4, 679-688 (2006).
5. MacDonald, M.L. et al. Identifying off-target effects and hidden phenotypes of drugs in human cells. Nat Chem Biol 2, 329-337 (2006).
6. Grasberger, B.L. et al. Discovery and cocrystal structure of benzodiazepinedione HDM2 antagonists that activate p53 in cells. J Med
Chem 48, 909-912 (2005).

You might also like