31
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- ing the measurement of catalytic activity from a purified enzyme 1 , a reconstituted complex 2,3 , a cellular extract 4,5 or a phenotype 6–9 in intact cells. Configuring assays to function within the constraints imposed by HTS differentiates an HTS assay from traditional laboratory assays, as outlined in Table 1. The 96-well microtiter plate, originally designed to facilitate serological studies in virology 10 , was later recognized as a suit- able replacement for cuvette, tube and dish-based assays. Combined with spectrophotometric plate readers, the microtiter plate quickly became the standard for performing parallel assays for rapid preliminary eval- uation of compound bioactivity. During the 1990s, the tremendous increase in chemical libraries and assays with sensitivities amenable to miniaturization further drove the parallel processing concept that today is practiced in the 96-, 384- or 1,536-microwell plate formats, among others. ‘High throughput’ is a relative term, but it is generally defined as the testing of 10,000 to 100,000 compounds per day, accomplished with mechanization that ranges from manually operated workstations to fully automated robotic systems 11,12 . Assays designed for the purpose of HTS attempt to integrate biological fidelity with enabling assay and screening technologies. Subsequently, understanding the pharmacological impact of the assay design is critical to the identification of biologically relevant compounds from HTS. Assay design for HTS An effective HTS strategy considers both the primary and subsequent secondary assay designs carefully. First, is the nature of the response to be measured clearly defined: does the signal increase, decrease, change in nature or location or belong to part of a more complex response? Can the response be measured via a single parameter, or is a multiparametric output possible? For example parameters that provide counter-screen information can guide active compound selection by differentiating selectivity among related targets 13 , or by distinguishing the inhibition of a cellular pathway from a cytotoxic response in a cell-based assay 14–16 . Second, is the stimulus dependent only on the compound being tested, or is the response to the compound conditioned by another stimulus, such as a fully efficacious concentration of agonist (when screening for an antagonist), a subefficacious concentration of agonist (when screen- ing for a potentiator) or a permissive temperature? In conditional assays, compound activities independent of the conditional stimulus can be false positives or can act outside the pathway of interest (Fig. 1d). Third, does the duration of the response occur within seconds upon application of the stimulus, thereby requiring a rapid read (for example, a calcium transient), or within minutes or hours of the stimulus, thereby permit- ting endpoint or multiple time point measurements? These parameters will define the basic format of the primary HTS. Follow-up assays, also referred to as ‘secondary screens’ or ‘counter screens’, should be planned before the HTS. It is essential to view the primary HTS as the initial step of an integrated process. Follow-up tests validate compounds targeting the intended biological interaction and assist in the elimination of compounds that generate a positive signal via other mechanisms. False positives can be anticipated based on the assay design (for example, fluorescent compounds in an assay with a fluorescent readout) or eliminated by testing in an independent or ‘orthogonal’ assay that uses a different methodology or biological read- out of target activity. As an example, cyclic AMP (cAMP), an important cellular second messenger, can be measured with an output response signal either directly or indirectly proportional to cellular [cAMP] (Fig. 2). cAMP assays showing inversely proportional responses to changes in [cAMP] in principle can function in an HTS-confirmatory assay sequence in which selected candidates identified in the HTS are then tested for initial validation of activity. Such ‘preliminary confirma- tion’ strategies using HTS assays in sequence (or parallel) can greatly lower the retesting of false positives in substantially lower throughput secondary assays that are generally more complex and designed to be a more physiological measure of the biological system under study 17 . US National Institutes of Health Chemical Genomics Center, National Institutes of Health, 9800 Medical Center Drive, Bethesda, Maryland 20892-3370, USA. Correspondence should be addressed to J.I. ([email protected]). Published online 18 July 2007; doi:10.1038/nchembio.2007.17 466 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY REVIEW

Review Inglese NatChemBiol 2007 Assays SuppTable

Embed Size (px)

Citation preview

Page 1: Review Inglese NatChemBiol 2007 Assays SuppTable

High-throughput screening assays for the identification of chemical probesJames 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-ing the measurement of catalytic activity from a purified enzyme1, a reconstituted complex2,3, a cellular extract4,5 or a phenotype6–9 in intact cells. Configuring assays to function within the constraints imposed by HTS differentiates an HTS assay from traditional laboratory assays, as outlined in Table 1. The 96-well microtiter plate, originally designed to facilitate serological studies in virology10, was later recognized as a suit-able replacement for cuvette, tube and dish-based assays. Combined with spectrophotometric plate readers, the microtiter plate quickly became the standard for performing parallel assays for rapid preliminary eval-uation of compound bioactivity. During the 1990s, the tremendous increase in chemical libraries and assays with sensitivities amenable to miniaturization further drove the parallel processing concept that today is practiced in the 96-, 384- or 1,536-microwell plate formats, among others. ‘High throughput’ is a relative term, but it is generally defined as the testing of 10,000 to 100,000 compounds per day, accomplished with mechanization that ranges from manually operated workstations to fully automated robotic systems11,12. Assays designed for the purpose of HTS attempt to integrate biological fidelity with enabling assay and screening technologies. Subsequently, understanding the pharmacological impact of the assay design is critical to the identification of biologically relevant compounds from HTS.

Assay design for HTSAn effective HTS strategy considers both the primary and subsequent secondary assay designs carefully. First, is the nature of the response to be measured clearly defined: does the signal increase, decrease, change in nature or location or belong to part of a more complex response? Can the response be measured via a single parameter, or is a multiparametric output possible? For example parameters that provide counter-screen

information can guide active compound selection by differentiating selectivity among related targets13, or by distinguishing the inhibition of a cellular pathway from a cytotoxic response in a cell-based assay14–16. Second, is the stimulus dependent only on the compound being tested, or is the response to the compound conditioned by another stimulus, such as a fully efficacious concentration of agonist (when screening for an antagonist), a subefficacious concentration of agonist (when screen-ing for a potentiator) or a permissive temperature? In conditional assays, compound activities independent of the conditional stimulus can be false positives or can act outside the pathway of interest (Fig. 1d). Third, does the duration of the response occur within seconds upon application of the stimulus, thereby requiring a rapid read (for example, a calcium transient), or within minutes or hours of the stimulus, thereby permit-ting endpoint or multiple time point measurements? These parameters will define the basic format of the primary HTS.

Follow-up assays, also referred to as ‘secondary screens’ or ‘counter screens’, should be planned before the HTS. It is essential to view the primary HTS as the initial step of an integrated process. Follow-up tests validate compounds targeting the intended biological interaction and assist in the elimination of compounds that generate a positive signal via other mechanisms. False positives can be anticipated based on the assay design (for example, fluorescent compounds in an assay with a fluorescent readout) or eliminated by testing in an independent or ‘orthogonal’ assay that uses a different methodology or biological read-out of target activity. As an example, cyclic AMP (cAMP), an important cellular second messenger, can be measured with an output response signal either directly or indirectly proportional to cellular [cAMP] (Fig. 2). cAMP assays showing inversely proportional responses to changes in [cAMP] in principle can function in an HTS-confirmatory assay sequence in which selected candidates identified in the HTS are then tested for initial validation of activity. Such ‘preliminary confirma-tion’ strategies using HTS assays in sequence (or parallel) can greatly lower the retesting of false positives in substantially lower throughput secondary assays that are generally more complex and designed to be a more physiological measure of the biological system under study17.

US National Institutes of Health Chemical Genomics Center, National Institutes of Health, 9800 Medical Center Drive, Bethesda, Maryland 20892-3370, USA. Correspondence should be addressed to J.I. ([email protected]).

Published online 18 July 2007; doi:10.1038/nchembio.2007.17

466 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY

R E V I E W

Page 2: Review Inglese NatChemBiol 2007 Assays SuppTable

Regardless of the approach taken, the HTS assay protocol must be simplified from its original form (Table 1), and the following param-eters must be optimized: (i) sufficient sensitivity to enable the identifi-cation of compounds with low potency or efficacy. (ii) Reproducibility and stability of the biological response between wells and plates. This parameter is dependent on both assay reagents and hardware (such as dispensers, incubators and detectors). (iii) Accuracy of the positive and negative controls as compared with the known pharmacology of the tar-get. (iv) Economic feasibility (frequently measured as cost per well).

Validation of an assay protocol determines its suitability for HTS and involves two parts: performance and sensitivity18 (Fig. 3). Several sta-tistical parameters are used to evaluate assay performance. Signal win-dow and Z factor (Table 2) are calculations that gauge the fold response between maximum and minimum output signals and the precision of this response within a plate and across plates. The signal window mea-sures the statistically significant difference between the maximum and minimum signal, but it is not as reliable as Z factor for predicting assay

performance18. The Z factor’s unitless scale (ranging from 0 to 1) allows comparison of different assays and screens using the control wells (Z') and sample wells (Z) of the plate. Comparison of Z' and Z in assay validation has been used to determine compound screening concentration, as assays in which Z' > Z pre-dict percent of actives greater than the <1% target value for single concentration screening paradigms. If compounds are absent from the sample field (for example, DMSO only), a low 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 validation is assay sensitivity—the accuracy and reproducibility of potency measurements of control compounds or conditions. The minimum significance ratio (MSR, Table 2) is used to measure the reproducibility of potency values and defines the statistically significant potency range that can be measured in an assay19,20. During HTS, the MSR can be calcu-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 HTSThe photon, which is created by either a fluo-rescent or luminescent mechanism, is the principal output in HTS assays. As the most widely used HTS detection modality (Table 3; see Supplementary Table 1 online for addi-tional detail), fluorescent energy is highly tunable, which allows the development of a variety of fluorescent sensors21. Fluorescence can be bright and occur on a timescale ranging from 10−9 s (for typical organic fluorophores) to 10−4 s (for lanthanide cryptates), thereby allowing for light’s many optical properties to be exploited by a number of detection methods22. For example, the application of plane-polarized light in fluorescence polar-

ization (FP) can be used to measure a probe’s rotational perturba-tions, thereby enabling simple assay designs dependent on a single labeled ligand (ref. 23, and ref. 24 and other articles in the same issue). Alternatively, time-resolvable fluorescent trivalent lanthanides such as Eu3+ and Tb3+ can be used to increase assay sensitivity by suppress-ing background fluorescence from assay components and library compounds25,26. Further, fluorescent probes can transfer energy to other chromophores or fluorophores in assays that measure distance-dependent fluorescence or fluorescence quenching via resonance energy transfer (FRET or QFRET) mechanisms25–27.

For cellular assays in particular, chemical28,29 and genetically encoded27 fluorescent probes allow customized, multiplexed or direct measurement of events from biological components present at low concentrations30. Coupled with scanning microplate cytometers31 or microscope-based imaging systems, these probes are moving HTS from population-aver-aged outputs toward the enumeration of individual cellular events8,9, thereby enhancing the information content obtainable from each well.

PEP ADP

or

GTP-γS

GTP-γS

Pyruvate ATP + luciferin

GPCR

hνEu

GFP

GFP

P

P

ab

dc

Figure 1 Assay categories and methodologies. (a) An isolated enzyme catalyzes turnover of a profluorescent blue- or red-emitting substrate. (b) A coupled-enzyme assay in which the ATP product, generated by the pyruvate kinase–catalyzed reaction (ribbon structure), is detected by a “reporter” enzyme (firefly luciferase) to create a luminescent output (hν). (c) A membrane preparation is used to measure the activity of a GPCR through the ligand-dependent activation of associated G proteins as measured by the binding of a fluorescent europium (Eu3+)-labeled GTP-γS analog. (d) A phenotypic 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 agent (, blue dashed arrow) as a prerequisite to revealing potentiators (compounds that are active only in the presence of the costimulus) of the pathway, as measured by the nuclear translocation of a GFP-labeled protein. Such potentiators could act by inhibiting proteins that repress the pathway (black dashed arrows).

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 467

R E V I E W

Page 3: Review Inglese NatChemBiol 2007 Assays SuppTable

The absence of excitation light energy in the generation of a lumi-nescence signal greatly reduces compound interference in HTS (Fig. 4). Although the signal strength can be significantly lower than fluores-cence, the background is negligible, which allows luminescence-based assays to have an enormous dynamic range, on the order of 1,000 times more sensitive than equivalent fluorescence-based assays32. Both che-miluminescence and bioluminescence use a chemical reaction to create light. In the case of bioluminescence, an enzyme (typically from the firefly Photinus pyrali) generates light through its action on a luciferin substrate, analogs of which have been used to develop assays for cyto-chrome P450s, proteases and monoamine oxidases32. Further, the active site of the P. pyrali luciferase has been mutated to spectrally vary the luminescence33, thereby allowing the use of multiple luminescent sen-sors in an assay15.

Luciferase reporters are prevalent in HTS assays because of convenient detection, ample sensitivity and their ubiquitous occurrence in academic research. Dual luciferase reporter assays with distinct substrate speci-ficities, kinetics or emission maxima have been useful for identifying activities specific to the signaling pathway of interest. Frequently, the second reporter is used to normalize gene expression or discriminate cytotoxic compounds34,35. Luciferase reporters can be combined with other detection formats as well, for instance in tandem with a green fluorescent protein (GFP) reporter14, β-galactosidase or Alamar blue16 to assess cytotoxicity.

The β-lactamase reporter offers particular advantages for HTS. For example, the β-lactam–coupled coumarin-fluorescein substrate, CCF4/AM, allows a ratiometric read that helps minimize variation in cell num-ber or substrate concentration because the emission maxima of the cleaved and intact β-lactamase substrates (460 and 530 nm, respectively) are dis-tinct (reviewed in ref. 36). In addition, the uncleaved substrate can be used to assess cytotoxicity, as fluorescence is detected only if the substrate is taken up by living cells and de-esterified37,38.

A hybrid system involving both bioluminescence and fluorescence is represented by BRET-based assays39 (Table 3 and Supplementary Table 1). Here luciferase luminescence stimulates a proximal fluorescent protein emission, thereby bypassing the need for exogenous excitation light. However, this technique may be more challenging to optimize for HTS assays because of lower S:B ratios and higher background from weakly associated target proteins40.

In addition, radioactive decay in conjunction with the scintillation proximity assay (SPA) format is applicable to a host of targets41, and low-cost absorbance assays have proven reliable in measuring analytes (for example, H+ or Pi), cellular staining or cell growth. However, in microtiter plates, absorbance measurements can have low S:B ratios

owing to the short path length that results from low volumes, though several assays have been implemented in 1,536-well plates42,43.

Variations on the use of these technologies in the design of assays are impressive, and a discussion of each is beyond the scope of this paper. Assays suitable for HTS comprise a subset of these and must meet the requirements outlined in Table 1. The next section highlights the well-exploited assay formats.

Assays for specific targetsHistorically, screening assays have been performed on a limited set of targets, many involving purified proteins. Approximately two thirds of therapeutic targets are comprised of enzymes and receptors44. More recently, HTS labs are increasing attention on cell-based high-content screens (HCS) that analyze biological events at subcellular resolution8,9; a 2006 survey indicated that roughly half of HTS assays are cell-based and that the use of HCS is rapidly growing, particularly in secondary screen-ing45. The types of proteins for which drugs have been successfully devel-oped are relatively restricted. About half of experimental and marketed drugs target five main protein families: G protein–coupled receptors (GPCRs), kinases, proteases, nuclear receptors (NRs) and ion channels46. Because of the great interest in these protein classes for drug develop-ment, many approaches and technologies (Table 3 and Supplementary Table 1) have been developed to assay these targets, and in some cases they are broadly applicable to other target classes as well.

Enzymes. Enzymes are comprised of six categories: oxidoreductases, transferases, hydrolases, lyases, isomerases and ligases. They catalyze a large number of chemical transformations. In cases in which the enzyme, cofac-tor and substrate can be obtained in an isolated, active form, the devel-opment of an HTS assay generally hinges on the availability of reagents and technologies to enable substrate or product detection. Here, protein kinases and proteases will serve to illustrate various approaches to HTS assay design for enzymes.

Protein kinases. Protein kinases mediate the most prevalent post-transla-tional modification in cells, phosphorylate an array of targets in signaling cascades and cycles, and are important for therapeutic intervention, par-ticularly oncology47. In this section we discuss biochemical approaches to protein kinase assays, as cell-based assays are covered in the later sections on cellular signaling. The biochemical approaches can be divided into two categories: generic assays independent of subfamily (for example, tyrosine versus serine/threonine) and antibody-based formats that detect an epitope within the phosphorylated product. The latter are generally used for a particular subfamily or specific protein kinase.

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, cell type

Compoundb, compound concentration

Assay container Varied—tube, slide, microtiter plate, Petri dish, cuvette, animal Microtiter plate

Time of measurement Milliseconds to months

Measurements as endpoint, multiple time points, or continuous

Minutes to hours

Measurements typically endpoint, but also pre-read and kinetic

Output formats Plate reader, radioactivity, size separation, object enumeration, images interpreted by human visual inspection

Plate reader—mostly fluorescence, luminescence and absorbance

Reporting format “Representative” data; statistical analysis of manually curated dataset Automated analysis of all data using statistical criteriaaSpecial reagent dispensers required. bIdeally available in milligram quantity with analytical verification of structure and purity.

468 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY

R E V I E W

Page 4: Review Inglese NatChemBiol 2007 Assays SuppTable

GPCR GPCRAC AC[cAMP]Gi Gs

EFC

TR-FRET

ED

D

D

ED-labeled cAMP Acceptor-labeled cAMP Cellular cAMP

EA

A

A

S

a

b

c

Act

ivity

(%

)A

ctiv

ity (

%)

Act

ivity

(%

)A

ctiv

ity (

%)

Act

ivity

(%

)A

ctiv

ity (

%)

100

75

50

25

0

100

75

50

25

0

100

75

50

25

0

100

75

50

25

0

100

75

50

25

0

100

75

50

25

0

Log[compound] (M)

Log[compound] (M)

Log[compound] (M)

Log[compound] (M) Log[compound] (M)

Log[compound] (M)

Gαi-coupled response Gαs-coupled response

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.

Protein kinases catalyze phosphoryl group transfers from ATP to hydroxyl groups of amino acid side chains of serine, threonine or tyrosine. An easily implemented assay format that has been used suc-cessfully in HTS detects ATP depletion via luciferase, an ATP-depen-dent photoenzyme48,49. A disadvantage of this approach is that 50% conversion of the ATP is required to achieve a two-fold S:B. This may require increased enzyme levels for sufficient signal, thus decreasing sensitivity and increasing reagent costs. Alternatively, kinase activity can be measured through ADP production. One assay format uses an ADP-specific antibody and labeled ADP to generate an FP competition binding assay to monitor ADP product formation50. A coupled enzyme strategy for ADP detection that yields a fluorescent readout has also been described51.

A number of methods measure kinase activity by the detection of phosphorylated peptide. Several formats use micron-sized beads or particles to partition the product phosphopeptide from ATP substrate or peptide. Radiometric filter binding remains the standard for kinase selectivity studies. For HTS, biotinylated 33P-phosphopeptide is typi-cally captured by streptavidin-coated SPA beads41 or scintillant-treated microtiter plates (for example, FlashPlates). In immobilized metal-ion affinity-based fluorescence polarization (IMAP), transition metal-che-lated particles bind fluorescently labeled phosphopeptides with high affinity, thereby enabling an antibody-independent FP format for the detection of phosphopeptide products52. Recently IMAP has been adapted to FRET-based systems in which size limitations on the poly-peptide substrate (typically <10,000 MW) are relaxed53. A similar system has been developed based on fluorescence quenching through an iron-phosphate interaction54. However, for HTS the ratiometric nature of IMAP is an advantage of this system over fluorescence quenching.

Antibody-based formats that detect a specific phosphorylated peptide or sequence capitalize on the many phosphospecific antibodies available and on nonseparation technologies, including TR-FRET, ALPHA and

FP. Such assays are advantageous in targeting kinase activity dependent on, for example, hierarchical phosphorylation or kinase cascades55. Also, multiplexed readouts of kinase activity and cellular phosphoproteins are possible with these formats56.

A unique system that assays the kinase inactive state relies on ATP binding and uses enzyme fragment complementation (EFC) for detec-tion57. Here the general kinase inhibitor, staurosporine, is linked to a enzyme donor (ED) fragment of β-galactosidase that when displaced from the kinase ATP pocket restores β-galactosidase activity through α complementation with the acceptor enzyme (EA) component58.

A final general point, illustrated with kinases, relates to assay designs that recapitulate as closely as possible the native target to uncover new mechanisms of compound action. As most biochemical kinase assays have used only the catalytic domain for HTS, ATP site–directed inhibitors are the primary class of agents identified. However, with a TR-FRET–based HTS of Akt-family kinases, Barnett et al.59 discovered a pharmaco-logically unique class of compounds using the full-length kinases. These compounds stabilized the inactive conformation of Akt1 and Akt2 via a site formed only in the presence of the pleckstrin homology (PH) domain of the kinase59, thereby allowing a unique mode of selectivity between Akt1 and Akt2 versus Akt3, which is not inhibited by the chemical class owing to its lack of a PH domain.

Proteases. As well-established drug targets60, proteases have received considerable coverage in terms of assay formats and reagent kits. Profluorescent or FRET-based substrates are commonly used to assay proteases as purified proteins or in cell lysates. Typically, probes are coupled to a two-to-five-amino-acid peptide comprising a protease cleavage site and become fluorescent after liberation by protease cleav-age13,61. Though ideally suited to HTS, a shortcoming of this approach is that the fluorogenic peptide is often insufficient in length to cover the entire binding pocket.

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 469

R E V I E W

Page 5: Review Inglese NatChemBiol 2007 Assays SuppTable

Although there are numerous ways to measure endoprotease activity, assays for exoproteases that recognize C- or N-terminal residues are far less available62. The hydrolytic activity of the protease has been used to process simple chromogenic esters and amides, but assays based on these substrates generally lack sensitivity for miniaturized HTS assays. One approach is to use protease-mediated “epitope unmasking,” in which cleavage by the exoprotease reveals a binding site for a detection reagent. Here, sensitive time-resolved FRET (TR-FRET)-based formats, for example, can be applied63.

Cellular FRET-based methods have been developed using fluorescent proteins or dyes linked by a protease consensus sequence. When expressed or taken up by cells, cleavage of the probe can be measured by a change in fluorescence upon activation of proteases such as caspase-3 (ref. 64) and hepatitis virus (HCV) NS2/3 (ref. 65). Similarly, cells expressing a fusion of GFP and alkaline phosphatase coupled through an NS3/4A cleavage site were used to monitor the activity of this HCV protease by the detection of secreted alkaline phosphatase66. In an approach based on cell viability, procaspase-3 was modified to contain a β-secretase cleavage motif that triggered apoptosis in 293T cells upon expression of β-secretase. β-secre-tase inhibitors were shown in this assay to promote cell survival67.

Nuclear receptors. NRs are transcription factors that are characterized by common modular features including a DNA-binding domain (DBD) and a ligand-binding domain (LBD). NRs are regulated by endobiotic ligands such as hormones and metabolites, or through xenobiotics68. The human genome encodes ~48 NR members of known and unknown function (orphan NRs) that are actively being studied in drug discovery and basic research. Consequently, the ligand or DNA-binding properties of isolated domains, and the cellular translocation and gene transactiva-tion properties of NRs, have been featured in assays designed for HTS (ref. 69 and other articles in the same focus issue).

Classical competition binding assays for NRs can be divided into radiometric and fluorometric assays. SPA formats have been developed for NRs such as peroxisome proliferator-activated receptor70, and FP has been used for well-characterized receptors such as the steroidal NRs, in which a fluorescent conjugate of a known receptor ligand and a purified LBD are prepared71.

Using GFP fusions of NRs8, cell imaging assays have been described that measure ligand-dependent, cytoplasm-to-nucleus translocation of the glucocorticoid receptor (GR)72 or engineered NR chimeras73. Translocation assays using GFP fusions of GR have been applied to 1,536-well systems (see ref. 8; PubChem BioAssay ID (AID) 450). Additionally, EFC assay formats have been described to measure NR translocation72,74 (PubChem AID 451). However, none of the aforemen-tioned assays can distinguish the functional effect of active compounds on NR transactivation.

To distinguish between agonist and antagonistic activity of com-pounds, the targeting of NRs has been addressed with TR-FRET ligand-dependent coregulator recruitment assays75, constructed with either coactivators or corepressors to identify agonists and antagonists. Additionally, cell-based reporter gene assays provide among the most sensitive means to screen for functional activity. These assays fuse NR response elements to reporters such as luciferase76, β-lactamase, and secreted alkaline phosphatase (ref. 69 and other articles in the same focus issue).

G protein–coupled receptors. GPCRs are targets of intensive drug development by the pharmaceutical industry, and numerous, well-estab-lished HTS assay methods exist77,78 and continue to be developed79. Technologies to measure ligand-receptor binding41,80 and the functional consequence of receptor activation77 have enabled screens that target the rich pharmacology of this receptor superfamily, including assays to differentiate full, partial and inverse agonists, antagonists and allosteric modulators81, and to identify ligands for orphan GPCRs8,82.

GPCRs broadly orchestrate a wide range of cellular events and pro-cesses for which HTS assays have been configured, including guanine nucleotide exchange5,83, modulation of second messenger concentra-tions84,85, ion channel activity86, protein phosphorylation56, protein trafficking8,87,88, pigment dispersion89, gene transcription6 and cell proliferation90, among others. The assays have different advantages and limitations, but collectively they illustrate the sophistication possible in HTS assay design given the current state of the art. For example, GPCRs that activate intracellular Ca2+ stores can be assayed using cal-cium-sensitive dyes (for example, Fluo-3 and Fura-4) and rapid-inject

Table 2 Equations for determining assay performance or sensitivity

Parameter Equation Comment

Coefficient of variation %CV = — × 100σµ

A measure of the precision relative to the mean value, calculated for the maximum and minimum signals. An acceptable limit is <15%.

Signal to noiseS:N = σmin

µmax – µmin A measure of the signal strength. This equation is sometimes given asa: µmax – µmin

σmax2 + σmin2Signal to background

S:B = µmin

µmax Usually calculated using control compounds, acceptable value >2-fold.

Signal windowSW = σmax

µmax – µmin – 3(σmax + σmin) The significant signal between max and min controls, acceptable value >2-foldb.

Z' factorZ' = 1–

µmax – µmin(3σmax + 3σmin) Representation of the SW using a score where a value >0.5 represents an acceptable

assay. Z' is measured in the absence of library compounds; Z is in the presence.

Minimum significance ratio MSR = 102σd Smallest potency ratio between two measurements that is statistically significant (with 95% confidence).

Cheng-PrusoffKi =

(1 + [S]/Km)IC50 Used to derive the Ki from the IC50 under conditions of competitive inhibition.

Cheng-Prusoff (ligand-binding) Ki =

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

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.

470 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY

R E V I E W

Page 6: Review Inglese NatChemBiol 2007 Assays SuppTable

imaging platforms84. However, GPCRs not naturally coupled to [Ca2+] modulation could not be assayed by this means until the development of so-called universal adaptors, molecular engineered G proteins91 or promiscuous G proteins92, to reconfigure such GPCRs to stimulate Ca2+ signaling. Today GPCR second messengers can be assayed by the direct measurement of cAMP85 (Fig. 2) or inositol phosphate species57,93.

Assays based on GPCR internalization are independent of G protein subtype, unlike second messenger formation, and have been applied suc-cessfully to analyze a number of GPCRs using either standard microtiter plate readers94 or fluorescence microscopy8,9. The use of the adaptor protein β-arrestin 2, fused with GFP, permits an assay design that gen-erally does not require modifications to the target receptor’s N or C terminus. Such phenotypic assays enabled with automated high-content image analysis are bridging cell biology with HTS in a way that will have a major impact on chemical biology, as discussed below.

Ion channels and transporters. Ion channels are membrane-spanning proteins that regulate electrochemical gradient-driven movement of inorganic ions such as Na+, K+, Ca2+ and Cl– into or out of cells. HTS technologies aimed at this target class are under rapid development95. The existence of multiple responsive states, including closed, open and inactive, to which drugs respond differently, add to the pharmacological diversity and assay development challenges96.

Ligand binding, ion flux, and fluorescent probe-based assays are the established methods for ion channel HTS (Table 3 and Supplementary Table 1). Competition binding measures the affinity of the ligand-channel protein interaction but is dependent on the availability of radiolabeled ligands. As with GPCRs, binding assays do not detect the functional consequence of compound interactions with ion channels. For these reasons, cell-based functional assays are more relevant to ion channel screening at both the primary and secondary levels.

The calcium-sensing fluorophores Fluo-3, Fluo-4 and Fura-2 have been used extensively in HTS for voltage- and ligand-gated channels and for GPCR-induced intracellular calcium release95. The fluorescence intensity of these probes inside cells increases proportionally with the elevation of intracellular free calcium concentration97. The addition of opaque dyes to the assay buffer eliminates the need for cell washing steps, as only intracellular fluorescence of adherent cells is detected in bottom-read measurements, thereby improving compatibility with HTS protocols95. Voltage-sensing probes track changes in membrane poten-tial and have been used in several assay formats, such as FRET-based and positional voltage sensors98.

Ion flux measurements using radioactive tracers such as 86Rb+, 45Ca2+, 22Na+ and 36Cl– are established assays of ion channel activity. However, the medium removal steps and considerable radioactivity limit this approach for HTS. Atomic absorbance spectrometry using nonradio-active tracer ions has recently permitted significant improvements to the ion flux assay for HTS99.

Patch-clamp electrophysiology remains the standard for measuring ion channel activity, as the potency of certain compounds, especially those with state- or voltage-dependent mechanisms, may appear signifi-cantly less potent when using other assay methods (for example, fluo-rometric or rubidium ion flux). Automated patch-clamp instruments represent a remarkable development in technology, increasing through-put from approximately ten compounds per week to hundreds95,100. However, further improvements are needed in seal quality, throughput, and cost before this technology can replace other formats such as fluo-rescence-based assays for HTS.

The broad examination of drug candidates for ion channel activity has become critical in the assessment of a candidate’s safety profile. In addi-tion to the functional inhibition of ion channels, the folding and traf-

ficking of ion channels represents another avenue for the development of HTS assays analogous to those developed for GPCRs. For example, an ELISA-based assay has been reported that measures the amount of cell surface hERG channel to determine drug effects on hERG channel trafficking101.

Transporters, another category of transmembrane proteins, have an important role in ion, membrane potential and nutrient homeostasis. Perturbations in transporter function can result in disease and altered drug absorption, distribution, metabolism, excretion and toxicity (ADMET) properties, leading to drug resistance, for example. As a case in point, multidrug-resistance (MDR) transporters (for example, P-gly-coprotein) belong to the family of ATP-binding cassette (ABC) proteins that extrude xenobiotic substances from the cell, including many drugs. Assays based on ATPase activity and fluorophore transport have been described102,103. Typical assay formats for amino acid and neurotransmit-ter transporters rely on radiolabeled substrates (for example, 3H-glycine) and involve multiple assay steps including medium aspirations. However, new assay methods are being developed that are suitable for HTS such as yeast heterologous expression systems104 and anion sensors105.

Assays for pathways and networks, and cellular phenotypesThe complexity of components and their interactions within cellular networks offers a rich source of new targets to assay using technologies to examine protein interactions, expression, distribution and stability. For instance, signaling mediated by Hedgehog, Wnt and Notch proteins uses unconventional proteins implicated in developmental and disease processes, making these pathways important therapeutic targets106 that are assayable with HTS automated fluorescence microscopy methods8. The following sections describe additional cellular assays but reveal only a glimpse of the breath of cellular processes amenable to testing by HTS assays.

Transcriptional readouts. Reporter gene expression is a mainstay for detection in cellular signaling assays6 and is useful for broadly identi-fying compounds that modulate the pathway and isolate new targets within. However, the purposely promiscuous nature of these assays is disadvantageous if the objective is a probe of a particular component of the pathway, as ‘target decryption’ can be a formidable undertaking. A separate critical consideration resides in identifying a cell type that appropriately reflects the cellular context of the signaling pathway being assayed by the reporter; both the types and amounts of cell surface recep-tors and intracellular signaling components can vary widely among the commonly used engineered cell lines.

Protein interactions and redistribution. Several innovative enzymatic and image-based methodologies have provided new ways to identify compounds that modulate movement and associations of protein com-ponents within living cells. These approaches fall into several general categories: reporter complementation, resonance energy transfer and protein redistribution. Low-affinity associations form the basis of pro-tein complementation assays (PCA), in which two proteins known to associate are separately fused to N- and C-terminal halves of a bifur-cated reporter protein. Upon interaction of the two sentinel proteins, the halves of the reporter (now in close proximity) reconstitute a func-tional enzymatic or fluorescent reporter. Commonly used PCA reporters include β-lactamase, GFP, YFP and luciferase27,40,107. This methodology is attractive because it can be adapted to many protein-protein interac-tions and is amenable to screening and profiling, as demonstrated with YFP-based108 and β-lactamase109 reporters. Alternatively, high-affinity β-galactosidase α complementation has been used to monitor protein movement. Here the ω fragment is targeted to a specific subcellular

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 471

R E V I E W

Page 7: Review Inglese NatChemBiol 2007 Assays SuppTable

region and the α-peptide-tagged translocating protein restores β-galac-tosidase activity once α and ω combine74.

In the resonance energy transfer methods FRET and BRET (Table 3 and Supplementary Table 1), donor and acceptor (or sensor) proteins are fused to the target proteins of interest. In BRET, a signal is generated when a bioluminescent donor enzyme is brought into proximity with an acceptor fluorophore, the photon emission of which is measured39. In FRET, energy is transferred from a donor to an acceptor fluorophore when the two proteins come into close proximity, which is manifested as a change in wavelength of the emitted fluorescence. FRET-based sensors use spectrally distinct fluorescent proteins such as engineered GFP and YFP (reviewed in ref. 27). Intracellular indicators of specific kinase activity have been described using phosphorylation-activated FRET sensors engineered from fluorescent protein pairs110, or for insulin signaling via PIP3 recruitment of labeled Akt2 PH domains111. Additionally, a cell-based HTS using BRET to detect CCR5-β arrestin association identified inhibitors of CCR5-mediated signaling94.

FRET, BRET, PCA and similar methods have been combined with subcellular imaging to produce highly successful and versatile HTS assays for monitoring protein movements within cells9,110. The develop-

ment of technology platforms for automated cellular imaging, together with sophisticated object recognition algorithms, has made the screen-ing of cell-based protein redistribution assays routine8,9. For instance, assays monitoring the redistribution of Akt1 (ref. 112), MAPK2 (ref. 113) and ERF1 (ref. 114) have been used for HTS. Indeed, MacDonald et al.108 report the use of 49 PCAs to monitor protein redistribution or interactions for profiling small-molecule activities.

Protein stability. The addition of degradation and destabilization domains to reporter proteins has been useful for assaying a number of cell processes (such as ubiquitin-mediated degradation and the cell cycle) and enabling screening of unstable proteins such as the cis-act-ing viral NS2/3 protease65. For pathways in which a particular protein is controlled through degradation, destabilization domain GFP sensors allow a measure of pathway activity. For instance, fusing the degrada-tion and localization domains of cyclin B1 onto GFP conferred onto the sensor the equivalent turnover and localization as endogenous cyclin B1, thereby allowing high-content screening for cell cycle modulators8. Similarly, fusion of the IκBα destabilization domain to luciferase was used to identify modulators of NFκB signaling15. Also, coupling of

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); 96-/384-well microtiter plates enabled with, for example, the ICR 12000.

99

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 oxygen reacts with acceptor bead within ~200 nm distance; λem 520–620 nm.

56,149

Bioluminescence Renilla or firefly luciferases used in reporter gene assays; analogs of D-luciferin used to expand biochemical assays for proteases, P450s, monoamine oxidases.

32

BRET Bioluminescence resonance energy transfer; useful for protein-protein interactions; Renilla luciferase donor luminescence (480 nm) stimulates fluorescence of GFP/YFP acceptor; follows dipole-dipole interaction 1/r6 (~5 nm maximum).

39,150

Bla/CCF4 FRET-based assay using β-lactamase and β-lactam-containing substrate (for example, CCF4); λex 395 nm, λem 460/530 nm; ratiometric reporter gene assays and other formats.

36,38,151

DELFIA Dissociation-enhanced lanthanide fluorescent immunoassay/time-resolved fluorescence; ELISA-like format using lanthanide chelate-labeled antibody; τf ~500 µs; separation-based enhanced fluorescence; variations using labeled ligands; extremely sensitive.

25

ECL Electrochemiluminescence assay. Uses microcarbon electrode plates allowing electrical oxidation of a [Ru(phen)3]2+ label. Multiplex applications in ELISA format.

17

EFC Enzyme fragment complementation. Based on β-galactosidase α complementation; high-affinity complementation. Applied to both cell-free and cell-based systems.

57,74

FLT Fluorescent lifetime; time-domain method (~ns) measures the rate of fluorescent decay following excitation; promises to eliminate fluorescence compound interference; only recently applied to some microplate systems.

146

FP Fluorescence polarization; polarized light reports change in probe ligand τc; fluorophores of τf ~ 5 ns needed to measure ligand of MW < 1,500 binding to a receptor of MW ≥ 10,000; widely applied, ratiometric measurement.

23,24

FRET Fluorescence resonance energy transfer; donor-acceptor pair as in EDANS-DABCYL peptides or fluorescent proteins (for example, CFP/YFP). Signal distance dependence ~5 nm maximum.

22,53

Fluorescence Measurement of fluorescence intensity including the use of chromophores for quenching as in profluorescent protease 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.

22,152

Fluorescence (rapid kinetic)

Requires fluorescent imaging plate reader or equivalent (for example, FLIPR, FDS6000, VIPR); timescale ~seconds; GPCR functional assays using Ca2+-sensitive dyes, or steady state ion channel activity using voltage-sensitive dyes.

84,98

GFP and variants Cell-based imaging microscopy/cytometry assays; translocation, redistribution, protein-protein interactions, reporter gene assays; see below.

27,112

HT electrophysiology High-throughput electrophysiology; measurement of ion channel currents with microaperture arrays such as the planar patch-clamp electrode (for example, PatchXpress 7000A, IonWorks Quattro).

98,99,153

PCA Protein complementation assay; recombinant split reporters (for example, GFP, luciferase, β-lactamase) fused to interacting proteins; phenotypic assays; low-affinity complementation.

40,108

SPA, FlashPlate Scintillation proximity assay; nonseparation-based radiometric detection using solid supports; typical isotopes include 35S, 33P or 3H.

41

TR-FRET Time-resolved FRET; also known commercially as HTRF/LANCE; nonseparation-based format using a lanthanide cryptate donor fluorophore of long fluorescence lifetime (τf ~ 500 µs); detection distance ~9 nm max.

25,26,53

τf, fluorescent lifetime; τc, rotational correlation time.

472 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY

R E V I E W

Page 8: Review Inglese NatChemBiol 2007 Assays SuppTable

the degradation domain of ornithine decarboxylase to the fluorescent protein ZsGreen was used to discover new proteosome inhibitors115. Fusing only the degradation domain of a particular protein, as opposed to the entire protein, to a sensor is advantageous in that it minimizes the pathway interference that can be caused by overexpression of the full-length protein8.

Assays for rare disease targets. With over 1,800 genetic associations with human disease now known116, the number of human diseases for which molecularly based assays could be developed is large. Occasionally a rare disease is caused by mutation of an immediately druggable tar-get, and even more rarely one that may be ameliorated by a clinically approved drug117. More commonly however, the genes responsible for rare diseases have obscure or entirely unknown functions. Some com-mon causes of human genetic diseases are listed below and illustrate the challenge of assay development for HTS in this arena:

(i) Nonsense mutations resulting in premature termination and absence of protein expression (for example, phenylketonuria, spinal muscular atrophy (SMA) and cystic fibrosis): assays to identify com-pounds that compensate for the defect, for example by upregulating an alternative pathway or pseudogene as in SMA118, or promoting readthrough of premature stop codons119.

(ii) Missense mutations resulting in a potentially functional but mis-folded or mistrafficked protein (for example, cystic fibrosis): assay needs to discover compounds that act as molecular chaperones to facilitate correct folding and trafficking of the protein to its proper location120.

(iii) Synonymous mutation that introduces a new splice donor or acceptor site, thereby altering splicing and producing an abnormal protein (for example, Hutchinson-Gilford Progeria Syndrome and β-thalassemia): assays to identify compounds that alter splicing121.

(iv) Motif expansion or duplication mutations that result in expres-sion of a protein with toxic or dominant negative function (for example, Huntington’s disease): assay needs to identify molecules that downregu-late the expression of the toxic protein or interfere with its function122.

These urgent and challenging biological problems will be an impor-tant frontier for HTS assay development in the next decade.

Assay optimization for HTSA balance between statistical performance and assay sensitivity facili-tates the identification of the broadest range of compound potencies and efficacies from HTS. Assays are generally configured to measure one or more of the following changes: (i) concentration of a substance (for example, enzyme substrate or product, second messenger or reporter gene product); (ii) concentration of a receptor-ligand complex or (iii) distribution of cellular markers. In addition, different methodologies measuring the same biological process can have discordant determi-nations of apparent compound potency and efficacy37,123–125, which emphasizes the need for selecting the proper assay format.

In general, four types of assays can be defined based on the acceptable limits of Z ' factor and assay sensitivity (Fig. 3a). Conditions leading to satisfactory performance (Z ' factor > 0.5) should be identified that minimally impact assay sensitivity of a control ligand or stimulus (Fig. 3a, quadrant i). With this in mind, the optimal assay will not always have the highest Z ' factor, as some very robust assays (Z ' scores > 0.8) can have poor sensitivity (Fig. 3a, quadrant iii). Conversely, assays with lower Z ' factors may have superior sensitivity (for example, Fig. 3a, quadrants i and ii) and may be rescued for HTS using replicate or concentration-dependent formats126. Because potent compounds can be identified in poorly designed assays, the irrational exuberance of observing such high-potency or high-efficacy compounds is often dis-placed by the realization that the activity was artifactual in nature, and

a careful analysis of the screening results ensues to identify compounds with reliable biological activity, many times with lower (µM) potency. Thus, the need for assay sensitivity is critically important to identify compounds with half-maximal inhibitory concentrations (IC50s) in the vicinity of the screening concentration, as these are false-negative candidates in assays having poor sensitivity.

Ideally, HTS assays are designed with substrate (S) or ligand (L) con-centrations below their Km or Kd, so that the observed potency (IC50) approximates the intrinsic potency Ki (see equations in Table 2; this relationship differs for more complex mechanisms or when determin-ing functional activity of receptors such as Kb values127–129). Though increasing either [S] or [L] improves the overall Z ' factor of the assay, doing so could potentially increase the IC50 values above the threshold used to identify actives in the screen (Fig. 3b,c), thereby preventing them from being identified as positives. Catalytic systems (for example, enzyme assays) should be designed for the minimum substrate con-version needed to achieve an acceptable Z ' factor. Although shifts in IC50s can be less than two-fold for enzyme assays operating at conver-sions upward of 80%130, high turnover significantly affects the ability to detect inhibitors with potencies near the screening concentration (Fig. 3d–f) and lowers the assay sensitivity for enzyme activators.

Many HTS technologies are inherently unable to detect low-affin-ity complexes (Kd ~ 1 µM) without resorting to high ligand or tar-get concentration, or engineering in a high-affinity interaction. The various assays developed to screen for inhibitors of the protein-protein interaction between the p53 tumor suppressor protein and its negative regulator (DM2) illustrate these issues (see Supplementary Table 2 online for assay details). In one assay the measurement of a p53-DM2 interaction with a competition-binding FP assay131 required 1 µM DM2 to allow sufficient complex formation (~50% bound) for a detectable signal. This configuration limits the observable compound potency range, thereby rendering the assay unreceptive to weak compounds and making it difficult to determine the rank order of potent compounds.

Conversely, high-affinity interactions resulting from enhanced avid-ity can increase the apparent affinity of the binding partners many fold through multiple receptor-ligand interactions. This “Velcro effect” is evident as a rightward shift in IC50 values of competing control ligands. For example, returning to the p53-DM2 interaction, Kane et al.132 developed a TR-FRET–based assay for this protein-protein interac-tion that required only nanomolar protein concentrations, in contrast to the micromolar level needed in the previous FP format. However, enhanced affinity due to multivalent interactions in the resulting com-plex (Supplementary Table 2) decreased the sensitivity of the assay as judged by the higher-than-expected IC50 of unlabeled competitor p53 protein. The potential for such avidity effects through surreptitious poly-valent interactions should always be considered carefully when examin-ing the benefits and limitations of the assay133. Further, irrespective of affinity, assay conditions requiring a high fraction of bound ligand (for example, most FP-based assays) cannot rely on the Cheng-Prusoff cor-rection to estimate the Kd or Ki of control or test compounds. Rather, accurate approximations can be made using equations describing the complete equilibria of the competing components134. In another bio-chemical format, the thermal shift assay135, which measures a ligand’s ability to stabilize a protein’s thermal denaturation, has been used to investigate compound binding to the p53 binding domain of DM2 directly. Though thermal shift technology at present is relatively low-throughput, requiring comparatively large amounts of target protein, it is one of the few function- or ligand-independent methods to screen for small molecule–protein interactions that may provide a means to identify small molecules with protein complex disruption potential (Supplementary Tables 1 and 2).

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 473

R E V I E W

Page 9: Review Inglese NatChemBiol 2007 Assays SuppTable

Owing to recent advances in high-throughput cellular imaging9,136, phenotypic assays now represent a viable alternative or complement to biochemical formats for protein-protein interaction HTS assays. Several cell-based assays have been used to investigate the p53-DM2 interac-tion108,137, and these may circumvent the affinity and ‘enhanced avidity’ concerns encountered in systems configured with isolated proteins.

Reagent-coated surfaces provide both opportunities and drawbacks for HTS assays. Formats involving reagent-coated beads or surfaces enable a physical separation of a bound and free ligand, thereby elimi-nating filtration steps41,133, or they provide components that comprise the detection system, as in ALPHA138. However, two issues should be considered when designing heterogeneous assays using solid supports. One issue pertains to the “avidity” of the interaction mentioned above that can occur in bead-based assays in which two binding partners are coated or tethered to separate beads, which results in multiple binding sites on the solid support (for example, the Velcro effect). The other issue is the capacity of the solid support to ‘hold’ a reagent or substrate. For example, when presenting an enzyme substrate on a solid support the amount of substrate that can be coated onto a surface can limit the reaction rate, and is often compensated by increasing enzyme concen-

tration to achieve a suitable signal139. The net result of increasing target concentration is to increase assay cost and decrease sensitivity140.

Compound interference within HTS assaysCompound-dependent assay interference is a very common cause of screening artifacts and must be ruled out before any compound activity is attributed to modulating the biological activity of interest. Here the inhibition or activation of an assay signal is not caused by the compound’s effect on the target or pathway of interest, but rather it is due to the compound’s ability to mimic that effect as a result of its physicochemical properties (for example, light absorbance) or adventi-tious biological activities (for example, luciferase inhibition)141,142 (Fig. 4). Cheminformatic approaches have been described that attempt to identify or remove from screening collections compounds with such problematic characteristics143, but currently prediction of such prop-erties is not reliable, and each active compound must be assumed to be an artifact until proven otherwise. This may be particularly true for chemical genomics applications, as the allowed functionalities of a chemical probe are less strict than for a medicinal lead. Thus although the boundaries of a useful chemical series may be broader for chemical

a b

d e f

cZ

' fac

tor

Z' factor

Z' factor

Z' factor Z' factor

Z' factor Z' factor

Inhi

bitio

n (%

)

Inhi

bitio

n (%

)

Conversion (%) Conversion (%)

Fol

d (I

C50

´/IC

50)

Fold (IC50´/IC50)

Fol

d (I

C50

/Ki)

Fo l

d (I

C50

´/IC

50)

100

75

50

25

0

100

75

50

25

0

12

10

8

6

4

2

0

1.0

0.8

0.6

0.4

0.2

0.0

1.0

0.8

0.6

0.4

0.2

0.0

1.0

0.8

0.6

0.4

0.2

0.0

12

10

8

6

4

2

0

12

10

8

6

4

2

0

Log[inhibitor] (M)

Log[inhibitor] (M)

Fraction bound receptor

i iii

ivii

FoldFold

Fold

1

0.5

03

0 25 50 75 100 0 25 50 75 100

0.2 0.4 0.6 0.8 1.0–6

–6–7–8–9

–5

–5

–4

–4

10

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, IC50 = 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).

474 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY

R E V I E W

Page 10: Review Inglese NatChemBiol 2007 Assays SuppTable

genomics applications than for drug development, the potential for misleading HTS activity is perhaps even greater with ‘probe-like’ than with ‘drug-like’ compounds.

Compound fluorescence and absorbance are very common causes of signal amplification or attenuation. Heterocyclic compounds that dominate screening collections most commonly fluoresce in the blue-green range, meaning that common assay sensors, such as GFP and coumarin, as well as the many profluorescent enzyme substrates that excite or emit near the blue end of the spectrum, are susceptible to this type of optical interference. Assay responses can differ in

signal strength or brightness and can determine assay sensitivity to certain classes of interferences, such as attenuation of luminescence or scintillation by light-absorbing compounds142. Signal brightness becomes particularly important with regard to compound interfer-ence when incident light is used to excite a fluorophore. In this case, assays having emitted light signals that overlap with the fluorescence from compounds will show increased false positives (Fig. 4a,b). The amount of compound fluorescence is dependent on the wavelength of excitation energy used, with blue light being more problematic than red light for typical heterocyclic compound libraries142,144.

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.

hν hν

Positives False positives

Emissionlight

Excitation light

a b c

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 475

R E V I E W

Page 11: Review Inglese NatChemBiol 2007 Assays SuppTable

To provide experimental data on the fluorescent properties of the compounds screened in our center, we recently performed spectro-scopic profiling of ~70,000 compounds from the Molecular Libraries Small Molecule Repository for eight common wavelengths (unpub-lished data; PubChem AIDs 587–592). For example, 4-methylumbel-liferone (4-MU; excitation 385 nm, emission 502 nm) is a commonly used label in profluorescent substrates for a wide range of hydrolytic enzymes. We found ~2% of the library fluoresced at a level equivalent to at least 100 nM 4-MU, which indicates that a screen using this excitation wavelength would yield many false positives via this mechanism.

Several approaches can minimize interference from fluorescent compounds, and should be used whenever possible. Separation-based formats such as ELISA offer an advantage over solution-phase assays, as compounds are removed before the detection step. However, the necessary separation steps can make ELISA difficult to miniaturize. For homogeneous assays, compound fluorescence can be reduced by using probes that use red-shifted fluorophores61, time-resolved lanthanide-based fluorescence25,26, pre-reads or multiple time points that permit background subtraction or rate determinations145. Technologies that in principle are minimally affected by compound fluorescence, such as fluorescence lifetime (FLT) spectroscopy146, are in development, but for now remain unproven in HTS.

Signal attenuation or activation of assay signal can also be caused by compound aggregation, which can result in target denaturation, ligand sequestration or light scattering147 (Fig. 4c,d). Like the optical properties discussed above, aggregation often occurs within the concen-tration range of the expected biological response and so may be easily mistaken for true modulation of the target biology. Aggregation effects are most prevalent at concentrations >1 µM, the hallmark of which is the elimination of apparent compound activity by the addition of deter-gents such as Triton X-100, a disruptor of aggregate micellar structures. Such compounds may show steep Hill slopes, but this alone does not distinguish aggregation from biologically relevant, stoichiometric, or cooperative modulation of enzyme activity. The prevalence of aggrega-tion-dependent enzyme inhibition has been observed to range from 2 to 10%, depending on the concentration of detergent and the chemi-cal library tested147; importantly, no clear structure-activity relation-ship has been evident from these studies, making the establishment of aggregation behavior currently entirely empirical. Results of a large aggregation screen recently performed at our center can be found in PubChem (AIDs 585 and 584).

Biological interferences are the second major source of compound-specific assay artifacts (Fig. 4a). These are due to inhibition (or, less frequently, activation) of a detection component of the assay such as inhibitors of enzyme reporters (for example, luciferases or β-lactamases) or intermolecular interactions (for example, nickel chelate–polyhisti-dine tag, small-molecule antigen-antibody, biotin-streptavidin). As an example, inhibition of luciferase by library compounds is common and has been described for the well-publicized compound resveratrol148. A recent screen of P. pyrali luciferase showed between 2 and 3% of the library was capable of interfering with luciferase detection (unpub-lished data; PubChem AID 411).

SummaryGiven the right compound library, assay design becomes the determining factor for identifying chemical probes using HTS. Before the selection of an assay format for HTS, thorough consideration of its advantages, limitations, and (when possible) prior application in HTS is recom-mended. The investigator’s familiarity with specific technologies should not form the basis of an assay strategy; rather, an unbiased assessment of key factors discussed here will increase the probability of a successful

HTS assay. Assay sensitivity and susceptibility to compound interference must also be studied prospectively when designing assays for HTS and planning follow-up analysis of active samples.

In general, the signal output of an assay should be assessed in con-junction with consequential controls. For example, protein-protein interaction assays using purified proteins may have suitable statistical performance factors but show poor sensitivity, which suggests that cel-lular assays with phenotypic outputs should be considered as alterna-tives. When investigating novel target biology, effort should be made to leverage existing well-developed technologies where possible. For instance, methods developed to detect ATP and ADP in protein kinase assays48,51,53 are also useful for lipid49 and metabolic kinases126.

As HTS assay technologies, screening systems, and analytical instru-mentation such as patch clamping and mass spectrometry evolve, the interfacing of large compound libraries with sophisticated assay and detection platforms will greatly expand the capability to identify chemi-cal probes for the vast untapped biology encoded by genomes. The rami-fications will allow a reinterpretation of the ‘druggable genome’.

Note: Supplementary information is available on the Nature Chemical Biology website.

ACKNOWLEDGMENTSWe thank D. Leja and J. Fekces for illustrations. The authors are supported by the Molecular Libraries Initiative of the NIH Roadmap for Medical Research and the Intramural Research Program of the National Human Genome Research Institute, NIH.

COMPETING INTERESTS STATEMENTThe authors declare no competing financial interests.

Published online at http://www.nature.com/naturechemicalbiologyReprints and permissions information is available online at http://npg.nature.com/reprintsandpermissions

1. Zhang, R. et al. A continuous spectrophotometric assay for the hepatitis C virus serine protease. Anal. Biochem. 270, 268–275 (1999).

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-tors. Anal. Biochem. 268, 318–329 (1999).

3. Seville, M., West, A.B., Cull, M.G. & McHenry, C.S. Fluorometric assay for DNA polymerases and reverse transcriptase. Biotechniques 21, 664–672 (1996).

4. Verma, R. et al. Ubistatins inhibit proteasome-dependent degradation by binding the ubiquitin chain. Science 306, 117–120 (2004).

5. 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).

6. Dinger, M.C. & Beck-Sickinger, A.G. in Molecular Biology in Medicinal Chemistry (eds. Dingermann, T., Steinhiber, D. & Folkers, G.) 73–94 (Wiley-VCH Verlag GmbH & Co., Weinheim, Germany, 2004).

7. Hodder, P., Mull, R., Cassaday, J., Berry, K. & Strulovici, B. Miniaturization of intracellular calcium functional assays to 1536-well plate format using a fluoro-metric imaging plate reader. J. Biomol. Screen. 9, 417–426 (2004).

8. Inglese, J. (ed.) Measuring Biological Responses with Automated Microscopy Vol. 414 (Elsevier Academic, San Diego, 2006).

9. Taylor, D.L., Haskins, J.R. & Giuliano, K.A. (eds.) High Content Screening Vol. 356 (Humana, Totowa, New Jersey, 2006).

10. Sever, J.L. Application of a microtechnique to viral serological investigations. J. Immunol. 88, 320–329 (1962).

11. Carroll, S.S., Inglese, J., Mao, S.-S. & Olsen, D.B. in Molecular Cancer Therapeutics: Strategies for Drug Discovery and Development (ed. Prendergast, G.C.) 119–140 (John Wiley & Sons, Inc., Hoboken, New Jersey, USA, 2004).

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-resolved quenching assay (TruPoint) for caspases 1, 3, and 6. Anal. Biochem. 325, 317–325 (2004).

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. 11, 469–480 (2006).

15. Davis, R.E. et al. A cell-based assay for IkappaBalpha stabilization using a two-color dual luciferase-based sensor. Assay Drug Dev. Technol. 5, 85–104 (2007).

16. O’Boyle, D.R. II et al. Development of a cell-based high-throughput specificity screen using a hepatitis C virus-bovine viral diarrhea virus dual replicon assay.

476 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY

R E V I E W

Page 12: Review Inglese NatChemBiol 2007 Assays SuppTable

Antimicrob. Agents Chemother. 49, 1346–1353 (2005).17. 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).

18. Iversen, P.W., Eastwood, B.J., Sittampalam, G.S. & Cox, K.L. A comparison of assay performance measures in screening assays: signal window, Z’ factor, and assay variability ratio. J. Biomol. Screen. 11, 247–252 (2006).

19. Eastwood, B.J. et al. The minimum significant ratio: a statistical parameter to characterize the reproducibility of potency estimates from concentration-response assays and estimation by replicate-experiment studies. J. Biomol. Screen. 11, 253–261 (2006).

20. Assay Guidance Manual Committee. Assay Guidance Manual Version 4.1 (Eli Lilly and Company and NIH Chemical Genomics Center, Bethesda, Maryland, 2005).

21. Fluorescence Spectra Viewer (Invitrogen Corporation, Carlsbad, California, 2006). <http://probes.invitrogen.com/servlets/spectraviewer?fileid1=1304dna>

22. Lakowicz, J.R. Principles of Fluorescence Spectroscopy (Springer, Berlin, 2006).

23. Owicki, J.C. Fluorescence polarization and anisotropy in high throughput screen-ing: perspectives and primer. J. Biomol. Screen. 5, 297–306 (2000).

24. Jameson, D.M. & Croney, J.C. Fluorescence polarization: past, present and future. Comb. Chem. High Throughput Screen. 6, 167–176 (2003).

25. Hemmila, I. & Laitala, V. Progress in lanthanides as luminescent probes. J. Fluoresc. 15, 529–542 (2005).

26. Trinquet, E. & Mathis, G. Fluorescence technologies for the investigation of chemical libraries. Mol. Biosyst. 2, 380–387 (2006).

27. 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).

28. Haugland, R.P. The Handbook—A Guide to Fluorescent Probes and Labeling Technologies 10th edn. (Eugene, Oregon, 2006).

29. Medintz, I.L., Uyeda, H.T., Goldman, E.R. & Mattoussi, H. Quantum dot biocon-jugates for imaging, labelling and sensing. Nat. Mater. 4, 435–446 (2005).

30. Haney, S.A., LaPan, P., Pan, J. & Zhang, J. High-content screening moves to the front of the line. Drug Discov. Today 11, 889–894 (2006).

31. 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).

32. Fan, F. & Wood, K.V. Bioluminescent assays for high-throughput screening. Assay Drug Dev. Technol. 5, 127–136 (2007).

33. Nakatsu, T. et al. Structural basis for the spectral difference in luciferase biolu-minescence. Nature 440, 372–376 (2006).

34. Blair, W.S. et al. A novel HIV-1 antiviral high throughput screening approach for the discovery of HIV-1 inhibitors. Antiviral Res. 65, 107–116 (2005).

35. 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).

36. Zlokarnik, G. Fusions to beta-lactamase as a reporter for gene expression in live mammalian cells. Methods Enzymol. 326, 221–244 (2000).

37. 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).

38. 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).

39. Pfleger, K.D. & Eidne, K.A. Illuminating insights into protein-protein interac-tions using bioluminescence resonance energy transfer (BRET). Nat. Methods 3, 165–174 (2006).

40. Remy, I. & Michnick, S.W. Mapping biochemical networks with protein-fragment complementation assays. Methods Mol. Biol. 261, 411–426 (2004).

41. Wu, S. & Liu, B. Application of scintillation proximity assay in drug discovery. BioDrugs 19, 383–392 (2005).

42. Lavery, P., Brown, M.J. & Pope, A.J. Simple absorbance-based assays for ultra-high throughput screening. J. Biomol. Screen. 6, 3–9 (2001).

43. Zuck, P. et al. Miniaturization of absorbance assays using the fluorescent proper-ties of white microplates. Anal. Biochem. 342, 254–259 (2005).

44. Zheng, C.J. et al. Therapeutic targets: progress of their exploration and investiga-tion of their characteristics. Pharmacol. Rev. 58, 259–279 (2006).

45. Fox, S. et al. High-throughput screening: update on practices and success. J. Biomol. Screen. 11, 864–869 (2006).

46. Hopkins, A.L. & Groom, C.R. The druggable genome. Nat. Rev. Drug Discov. 1, 727–730 (2002).

47. Pawson, T. & Scott, J.D. Protein phosphorylation in signaling—50 years and counting. Trends Biochem. Sci. 30, 286–290 (2005).

48. Singh, P., Harden, B.J., Lillywhite, B.J. & Broad, P.M. Identification of kinase inhibitors by an ATP depletion method. Assay Drug Dev. Technol. 2, 161–169 (2004).

49. Munagala, N. et al. Identification of small molecule ceramide kinase inhibitors using a homogeneous chemiluminescence high throughput assay. Assay Drug Dev. Technol. 5, 65–73 (2007).

50. Lowery, R.G. & Kleman-Leyer, K. Transcreener: screening enzymes involved in covalent regulation. Expert Opin. Ther. Targets 10, 179–190 (2006).

51. Charter, N.W., Kauffman, L., Singh, R. & Eglen, R.M. A generic, homogenous method for measuring kinase and inhibitor activity via adenosine 5′-diphosphate accumulation. J. Biomol. Screen. 11, 390–399 (2006).

52. 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).

53. Comley, J. Kinase screening and profiling-spoilt for choice. Drug Discov. World 7, 27–50 (2006).

54. Morgan, A.G., McCauley, T.J., Stanaitis, M.L., Mathrubutham, M. & Millis, S.Z. Development and validation of a fluorescence technology for both primary and secondary screening of kinases that facilitates compound selectivity and site-spe-cific inhibitor determination. Assay Drug Dev. Technol. 2, 171–181 (2004).

55. Newbatt, Y. et al. Identification of inhibitors of the kinase activity of oncogenic V600E BRAF in an enzyme cascade high-throughput screen. J. Biomol. Screen. 11, 145–154 (2006).

56. 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).

57. 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).

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. Barnett, S.F. et al. Identification and characterization of pleckstrin-homology-domain-dependent and isoenzyme-specific Akt inhibitors. Biochem. J. 385, 399–408 (2005).

60. Leung, D., Abbenante, G. & Fairlie, D.P. Protease inhibitors: current status and future prospects. J. Med. Chem. 43, 305–341 (2000).

61. Grant, S.K., Sklar, J.G. & Cummings, R.T. Development of novel assays for proteo-lytic enzymes using rhodamine-based fluorogenic substrates. J. Biomol. Screen. 7, 531–540 (2002).

62. Auld, D. in Proteolytic Enzymes Tools: and Targets (eds. Stocker, W. & Sterchi, E. E.) 30–48 (Springer-Velag, Berlin Heidelberg, 1999).

63. Ferrer, M. et al. Miniaturizable homogenous time-resolved fluorescence assay for carboxypeptidase B activity. Anal. Biochem. 317, 94–98 (2003).

64. Jones, J., Heim, R., Hare, E., Stack, J. & Pollok, B.A. Development and applica-tion of a GFP-FRET intracellular caspase assay for drug screening. J. Biomol. Screen. 5, 307–317 (2000).

65. Whitney, M. et al. A collaborative screening program for the discovery of inhibi-tors of HCV NS2/3 cis-cleaving protease activity. J. Biomol. Screen. 7, 149–154 (2002).

66. Lee, J.C. et al. Development of a cell-based assay for monitoring specific hepa-titis C virus NS3/4A protease activity in mammalian cells. Anal. Biochem. 316, 162–170 (2003).

67. Li, Y., Liu, Y.W. & Cordell, B. Novel functional assay for proteases and modulators. Application in beta-secretase studies. Mol. Biotechnol. 18, 1–10 (2001).

68. Olefsky, J.M. Nuclear receptor minireview series. J. Biol. Chem. 276, 36863–36864 (2001).

69. Zeh, K. et al. Gain-of-function somatic cell lines for drug discovery applications generated by homologous recombination. Assay Drug Dev. Technol. 1, 755–765 (2003).

70. Sun, S., Almaden, J., Carlson, T.J., Barker, J. & Gehring, M.R. Assay develop-ment and data analysis of receptor-ligand binding based on scintillation proximity assay. Metab. Eng. 7, 38–44 (2005).

71. Parker, G.J., Law, T.L., Lenoch, F.J. & Bolger, R.E. Development of high through-put screening assays using fluorescence polarization: nuclear receptor-ligand-binding and kinase/phosphatase assays. J. Biomol. Screen. 5, 77–88 (2000).

72. Fung, P. et al. A homogeneous cell-based assay to measure nuclear translocation using beta-galactosidase enzyme fragment complementation. Assay Drug Dev. Technol. 4, 263–272 (2006).

73. Mackem, S., Baumann, C.T. & Hager, G.L. A glucocorticoid/retinoic acid receptor chimera that displays cytoplasmic/nuclear translocation in response to retinoic acid. A real time sensing assay for nuclear receptor ligands. J. Biol. Chem. 276, 45501–45504 (2001).

74. Wehrman, T.S., Casipit, C.L., Gewertz, N.M. & Blau, H.M. Enzymatic detection of protein translocation. Nat. Methods 2, 521–527 (2005).

75. Gowda, K., Marks, B.D., Zielinski, T.K. & Ozers, M.S. Development of a coactiva-tor displacement assay for the orphan receptor estrogen-related receptor-gamma using time-resolved fluorescence resonance energy transfer. Anal. Biochem. 357, 105–115 (2006).

76. Zhu, Z. et al. Correlation of high-throughput pregnane X receptor (PXR) transac-tivation and binding assays. J. Biomol. Screen. 9, 533–540 (2004).

77. Eglen, R.M., Bosse, R. & Reisine, T. Emerging concepts of guanine nucleo-tide-binding protein-coupled receptor (GPCR) function and implications for high throghput screening. Assay Drug Dev. Technol. 5, 425–451 (2007).

78. Jacoby, E., Bouhelal, R., Gerspacher, M. & Seuwen, K. The 7 TM G-protein-coupled receptor target family. ChemMedChem 1, 761–782 (2006).

79. 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).

80. Handl, H.L. & Gillies, R.J. Lanthanide-based luminescent assays for ligand-recep-tor interactions. Life Sci. 77, 361–371 (2005).

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 477

R E V I E W

Page 13: Review Inglese NatChemBiol 2007 Assays SuppTable

81. Kenakin, T. Inverse, protean, and ligand-selective agonism: matters of receptor conformation. FASEB J. 15, 598–611 (2001).

82. Behan, D.P. & Chalmers, D.T. The use of constitutively active receptors for drug discovery at the G protein-coupled receptor gene pool. Curr. Opin. Drug Discov. Devel. 4, 548–560 (2001).

83. Labrecque, J. et al. The development of an europium-GTP assay to quantitate che-mokine antagonist interactions for CXCR4 and CCR5. Assay Drug Dev. Technol. 3, 637–648 (2005).

84. Chambers, C. et al. Measuring intracellular calcium fluxes in high throughput mode. Comb. Chem. High Throughput Screen. 6, 355–362 (2003).

85. Williams, C. cAMP detection methods in HTS: selecting the best from the rest. Nat. Rev. Drug Discov. 3, 125–135 (2004).

86. Reinscheid, R.K., Kim, J., Zeng, J. & Civelli, O. High-throughput real-time moni-toring of Gs-coupled receptor activation in intact cells using cyclic nucleotide-gated channels. Eur. J. Pharmacol. 478, 27–34 (2003).

87. Adie, E.J. et al. CypHer 5: a generic approach for measuring the activation and trafficking of G protein-coupled receptors in live cells. Assay Drug Dev. Technol. 1, 251–259 (2003).

88. Charest, P.G., Terrillon, S. & Bouvier, M. Monitoring agonist-promoted confor-mational changes of beta-arrestin in living cells by intramolecular BRET. EMBO Rep. 6, 334–340 (2005).

89. Karlsson, A.M. et al. Biosensing of opioids using frog melanophores. Biosens. Bioelectron. 17, 331–335 (2002).

90. Burstein, E.S. et al. Integrative functional assays, chemical genomics and high throughput screening: harnessing signal transduction pathways to a common HTS readout. Curr. Pharm. Des. 12, 1717–1729 (2006).

91. Liu, A.M. et al. Galpha(16/z) chimeras efficiently link a wide range of G pro-tein-coupled receptors to calcium mobilization. J. Biomol. Screen. 8, 39–49 (2003).

92. Kostenis, E. Is Galpha16 the optimal tool for fishing ligands of orphan G-protein-coupled receptors? Trends Pharmacol. Sci. 22, 560–564 (2001).

93. 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).

94. 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).

95. Zheng, W., Spencer, R.H. & Kiss, L. High throughput assay technologies for ion channel drug discovery. Assay Drug Dev. Technol. 2, 543–552 (2004).

96. McDonough, S.I. & Bean, B.P. in Voltage-Gated Ion Channels as Drug Targets Vol. 29 (eds. Triggle, D.J., Gopalakrishnan, M., Rampe, D. & Zheng, W.) 19–35 (Wiley-VCH, Weinheim, Germany, 2006).

97. Gee, K.R. et al. Chemical and physiological characterization of fluo-4 Ca(2+)-indicator dyes. Cell Calcium 27, 97–106 (2000).

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. Terstappen, G.C. Nonradioactive rubidium ion efflux assay and its applications in drug discovery and development. Assay Drug Dev. Technol. 2, 553–559 (2004).

100. 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).

101. Wible, B.A. et al. HERG-Lite: a novel comprehensive high-throughput screen for drug-induced hERG risk. J. Pharmacol. Toxicol. Methods 52, 136–145 (2005).

102. Schwab, D., Fischer, H., Tabatabaei, A., Poli, S. & Huwyler, J. Comparison of in vitro P-glycoprotein screening assays: recommendations for their use in drug discovery. J. Med. Chem. 46, 1716–1725 (2003).

103. Henrich, C.J. et al. A high-throughput cell-based assay for inhibitors of ABCG2 activity. J. Biomol. Screen. 11, 176–183 (2006).

104. Wieczorke, R., Dlugai, S., Krampe, S. & Boles, E. Characterisation of mamma-lian GLUT glucose transporters in a heterologous yeast expression system. Cell. Physiol. Biochem. 13, 123–134 (2003).

105. Verkman, A.S. Drug discovery in academia. Am. J. Physiol. Cell Physiol. 286, C465–C474 (2004).

106. Rubin, L.L. & de Sauvage, F.J. Targeting the Hedgehog pathway in cancer. Nat. Rev. Drug Discov. 5, 1026–1033 (2006).

107. Massoud, T.F., Paulmurugan, R., De, A., Ray, P. & Gambhir, S.S. Reporter gene imaging of protein-protein interactions in living subjects. Curr. Opin. Biotechnol. 18, 31–37 (2007).

108. MacDonald, M.L. et al. Identifying off-target effects and hidden phenotypes of drugs in human cells. Nat. Chem. Biol. 2, 329–337 (2006).

109. Buttner, F.H., Kumpf, R., Menzel, S., Reulle, D. & Valler, M.J. Evaluation of the InteraX system technology in a high-throughput screening environment. J. Biomol. Screen. 10, 485–494 (2005).

110. Wallrabe, H. & Periasamy, A. Imaging protein molecules using FRET and FLIM microscopy. Curr. Opin. Biotechnol. 16, 19–27 (2005).

111. Marine, S. et al. A miniaturized cell-based fluorescence resonance energy transfer assay for insulin-receptor activation. Anal. Biochem. 355, 267–277 (2006).

112. 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).

113. Almholt, D.L. et al. Nuclear export inhibitors and kinase inhibitors identified using a MAPK-activated protein kinase 2 redistribution screen. Assay Drug Dev. Technol. 2, 7–20 (2004).

114. 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).

115. Rickardson, L., Wickstrom, M., Larsson, R. & Lovborg, H. Image-based screen-ing for the identification of novel proteasome inhibitors. J. Biomol. Screen. 12, 203–210 (2007).

116. Online Mendelian Inheritance of Man (OMIM). (Johns Hopkins University and National Center for Biotechnology Information, National Library of Medicine, 2007). <http://www.ncbi.nlm.nih.gov/omim/>

117. Habashi, J.P. et al. Losartan, an AT1 antagonist, prevents aortic aneurysm in a mouse model of Marfan syndrome. Science 312, 117–121 (2006).

118. Jarecki, J. et al. Diverse small-molecule modulators of SMN expression found by high-throughput compound screening: early leads towards a therapeutic for spinal muscular atrophy. Hum. Mol. Genet. 14, 2003–2018 (2005).

119. Welch, E.M. et al. PTC124 targets genetic disorders caused by nonsense muta-tions. Nature 447, 87–91 (2007).

120. Bernier, V., Lagace, M., Bichet, D.G. & Bouvier, M. Pharmacological chaperones: potential treatment for conformational diseases. Trends Endocrinol. Metab. 15, 222–228 (2004).

121. Sazani, P. et al. Nuclear antisense effects of neutral, anionic and cationic oligo-nucleotide analogs. Nucleic Acids Res. 29, 3965–3974 (2001).

122. Zhang, X. et al. A potent small molecule inhibits polyglutamine aggregation in Huntington’s disease neurons and suppresses neurodegeneration in vivo. Proc. Natl. Acad. Sci. USA 102, 892–897 (2005).

123. Kenakin, T. Predicting therapeutic value in the lead optimization phase of drug discovery. Nat. Rev. Drug Discov. 2, 429–438 (2003).

124. Van Der Hee, R.M., Deurholt, T., Gerhardt, C.C. & De Groene, E.M. Comparison of 3 AT1 receptor binding assays: filtration assay, ScreenReady Target, and WGA Flashplate. J. Biomol. Screen. 10, 118–126 (2005).

125. Wu, X., Glickman, J.F., Bowen, B.R. & Sills, M.A. Comparison of assay technolo-gies for a nuclear receptor assay screen reveals differences in the sets of identi-fied functional antagonists. J. Biomol. Screen. 8, 381–392 (2003).

126. Inglese, J. et al. Quantitative high-throughput screening: a titration-based approach that efficiently identifies biological activities in large chemical librar-ies. Proc. Natl. Acad. Sci. USA 103, 11473–11478 (2006).

127. Burlingham, B. & Widlanski, T. An intuitive look at the relationship of Ki and IC50: a more general use for the Dixon plot. J. Chem. Educ. 80, 214–218 (2003).

128. Copeland, R.A. Mechanistic considerations in high-throughput screening. Anal. Biochem. 320, 1–12 (2003).

129. Leff, P. & Dougall, I.G. Further concerns over Cheng-Prusoff analysis. Trends Pharmacol. Sci. 14, 110–112 (1993).

130. Wu, G., Yuan, Y. & Hodge, C.N. Determining appropriate substrate conversion for enzymatic assays in high-throughput screening. J. Biomol. Screen. 8, 694–700 (2003).

131. 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).

132. Kane, S.A. et al. Development of a binding assay for p53/HDM2 by using homo-geneous time-resolved fluorescence. Anal. Biochem. 278, 29–38 (2000).

133. Zuck, P. et al. Ligand-receptor binding measured by laser-scanning imaging. Proc. Natl. Acad. Sci. USA 96, 11122–11127 (1999).

134. Zhang, R. et al. Fluorescence polarization assay and inhibitor design for MDM2/p53 interaction. Anal. Biochem. 331, 138–146 (2004).

135. Parks, D.J. et al. 1,4-Benzodiazepine-2,5-diones as small molecule antagonists of the HDM2-p53 interaction: discovery and SAR. Bioorg. Med. Chem. Lett. 15, 765–770 (2005).

136. Inglese, J. Engaging science and technology for drug development. Assay Drug Dev. Technol. 1, 1 (2002).

137. 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 interac-tions. Assay Drug Dev. Technol. 4, 679–688 (2006).

138. Peppard, J. et al. Development of a high-throughput screening assay for inhibi-tors of aggrecan cleavage using luminescent oxygen channeling (AlphaScreen). J. Biomol. Screen. 8, 149–156 (2003).

139. DeForge, L.E. et al. Substrate capacity considerations in developing kinase assays. Assay Drug Dev. Technol. 2, 131–140 (2004).

140. Iyer, R., Barrese, A.A. III, Parakh, S., Parker, C.N. & Tripp, B.C. Inhibition profil-ing of human carbonic anhydrase II by high-throughput screening of structur-ally diverse, biologically active compounds. J. Biomol. Screen. 11, 782–791 (2006).

141. Comley, J. Assay Interference, a limiting factor in HTS? Drug Discov. World 4, 91–98 (2003).

142. 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).

143. Rishton, G.M. Nonleadlikeness and leadlikeness in biochemical screening. Drug Discov. Today 8, 86–96 (2003).

144. Swift, K., Anderson, S.N. & Matayoshi, E.D. in Advances in Fluorescence Sensing Technology V, Proceedings of SPIE - The International Society for Optical

478 VOLUME 3 NUMBER 8 AUGUST 2007 NATURE CHEMICAL BIOLOGY

R E V I E W

Page 14: Review Inglese NatChemBiol 2007 Assays SuppTable

Engineering 4252 47–58 (2001).145. Porcari, V., Magnoni, L., Terstappen, G.C. & Fecke, W. A continuous time-

resolved fluorescence assay for identification of BACE1 inhibitors. Assay Drug Dev. Technol. 3, 287–297 (2005).

146. 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).

147. Feng, B.Y. et al. A high-throughput screen for aggregation-based inhibition in a large compound library. J. Med. Chem. 50, 2385–2390 (2007).

148. Bakhtiarova, A. et al. Resveratrol inhibits firefly luciferase. Biochem. Biophys. Res. Commun. 351, 481–484 (2006).

149. Ullman, E.F. et al. Luminescent oxygen channeling immunoassay: measurement

of particle binding kinetics by chemiluminescence. Proc. Natl. Acad. Sci. USA 91, 5426–5430 (1994).

150. Issad, T., Blanquart, C. & Gonzalez-Yanes, C. The use of bioluminescence reso-nance energy transfer for the study of therapeutic targets: application to tyrosine kinase receptors. Expert Opin. Ther. Targets 11, 541–556 (2007).

151. Zlokarnik, G. et al. Quantitation of transcription and clonal selection of single living cells with beta-lactamase as reporter. Science 279, 84–88 (1998).

152. Kumaraswamy, S. et al. Fluorescent-conjugated polymer superquenching facili-tates highly sensitive detection of proteases. Proc. Natl. Acad. Sci. USA 101, 7511–7515 (2004).

153. Kiss, L. et al. High throughput ion-channel pharmacology: planar-array-based voltage clamp. Assay Drug Dev. Technol. 1, 127–135 (2003).

NATURE CHEMICAL BIOLOGY VOLUME 3 NUMBER 8 AUGUST 2007 479

R E V I E W

Page 15: Review Inglese NatChemBiol 2007 Assays SuppTable

Supplementary Table 1: Assay Technologies Additional Entries

Technology Description Refs. Technology

Refs. Assays

AAS Atomic absorption spectroscopy for measurements of ion transport/flux (primarily focused on K+ channels with Rb+ as tracer); 96/384-well microtiter plates enabled with, for example, the ICR 12000.

1, 2 3-7

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

~ 200 nm dia.;λex 680 nm (laser); singlet oxygen reacts with acceptor bead within ~200 nm distance; λem 520-620 nm

11, 12 13-16

Bioluminescence Renilla or firefly luciferases used in reporter gene assays; analogs of D-luciferin used to expand biochemical assays for proteases, P450s, monoamine oxidases

17 18-27

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.

28, 29 30-37

BRET Bioluminescence resonance energy transfer; useful for protein-protein interactions; Renilla luciferase donor luminescence (480 nm) stimulates fluorescence of GFP/YFP acceptor; follows dipole-dipole interaction 1/r6 (~ 5 nm maximum)

38, 39 40

DELFIA Dissociation-enhanced lanthanide fluorescent immunoassay/Time resolved fluorescence; ELISA-like format using lanthanide chelate-labeled antibody; τf ~ 500 usec; separation–based enhanced fluorescence; variations employing labeled ligands; extremely sensitive.

41, 42 43-49

ECL Electrochemiluminescence assay. Uses micro-carbon electrode plates allowing electrical oxidation of a Ru(phen)3

2+ label. Multiplex applications in ELISA format.

50 20

Page 16: Review Inglese NatChemBiol 2007 Assays SuppTable

EFC Enzyme fragment complementation. Based on β-galactosidase α-complementation; High affinity complementation. Applied to both cell-free and cell-based systems.

51, 52 53-58

FCS / FIDA Fluorescence correlation spectroscopy / Fluorescence intensity distribution 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.

59, 60 61-65

FLT Fluorescent lifetime; time-domain method (~ns) measures the rate of fluorescent decay following excitation; promises to eliminate fluorescence compound interference; only recently applied to some microplate systems.

66

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.

67, 68 69-78

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

79, 80 81-84

Fluorescence

Measurement of fluorescence intensity including the use of chromophores for 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)

80, 85 86-97

Fluorescence (rapid kinetic)

Requires fluorescent imaging plate reader or equivalent (e.g. FLIPR, FDS6000, VIPR); timescale ~ seconds; GPCR functional assays using Ca2+ sensitive dyes, or steady-state ion channel activity using voltage sensitive dyes.

98, 99 99-108

Page 17: Review Inglese NatChemBiol 2007 Assays SuppTable

GFP and variants Cell-based imaging microscopy/cytometry assays; translocation, redistribution, protein-protein interactions, reporter gene assays; HTS enabled with microplate-compatible microscope-based imagers (e.q., ArrayScan, Opera, Pathway BioImager, and INCell imagers) and cytometers (below).

109-112 18, 110,

113-118

HT Electrophysiology

High-throughput electrophysiology; measurement of ion channel currents with microaperture arrays such as the planar patch-clamp electrode (e.g., PatchXpress 7000A, IonWorks Quattro). Transporter assays using solid supported membranes (e.g., SURFE2R) (see also ion flux measurements via AAS)

119-122 123-127

“Label Free” See following three entries

Spectrometry Mass spectrometry (e.g., RapidFire and SpeedScreen), high-field NMR (SAR by NMR), fragment based methods using X-ray crystallography.

128-131 132-135

Optical Refractive index based. Measurement of ligand-binding using prism-coupled 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.

136-138 139, 140

Non-optical Includes measurements of electrical impedance, quartz crystal acoustic 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.

141-146 147-151

Laser Cytometry See following three entries

Acumen-enabled Commonly referred to as ‘Laser-scanning fluorescence microplate cytometry’ 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).

152 153

Page 18: Review Inglese NatChemBiol 2007 Assays SuppTable

FMAT-enabled Fluorometric microvolume assay technology; also known as, Laser-scanning 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).

154, 155 43, 156-162

Flow-based Flow cytometry detection using laser-induced fluorescence or light scatter from 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.

163-165 166-168

PCA Protein complementation assay; recombinant split reporters (e.g., GFP, luciferase, β-lactamase) fused to interacting proteins; phenotypic assays; low affinity complementation.

169, 170 153, 171,

172 SPA, FlashPlate Scintillation proximity assay; non-separation based radiometric detection using

solid supports; typical isotopes include 35S, 33P or 3H

173 174-186

ThermoFluor Temperature-dependent protein unfolding is monitored using environmentally 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.

187 188

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

41, 42, 79 44, 189-201

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

Page 19: Review Inglese NatChemBiol 2007 Assays SuppTable

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).

Page 20: Review Inglese NatChemBiol 2007 Assays SuppTable

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).

Page 21: Review Inglese NatChemBiol 2007 Assays SuppTable

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).

Page 22: Review Inglese NatChemBiol 2007 Assays SuppTable

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).

Page 23: Review Inglese NatChemBiol 2007 Assays SuppTable

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).

Page 24: Review Inglese NatChemBiol 2007 Assays SuppTable

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).

Page 25: Review Inglese NatChemBiol 2007 Assays SuppTable

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).

Page 26: Review Inglese NatChemBiol 2007 Assays SuppTable

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).

Page 27: Review Inglese NatChemBiol 2007 Assays SuppTable

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).

Page 28: Review Inglese NatChemBiol 2007 Assays SuppTable

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).

Page 29: Review Inglese NatChemBiol 2007 Assays SuppTable

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).

Page 30: Review Inglese NatChemBiol 2007 Assays SuppTable

Components Format Technology p53 DM2

Conditions / Comment Ref:

Competition Binding

FP FITC-[15-30]p53 peptidea (Kd, app =2 uM); low affinity peptide FITC-11, 11 amino acid phage peptide (Kd, app =1 uM)

GST-[9-149]DM2 • 15 nM FITC peptide • 1 uM DM2 • 384-well plates • Non-Langmuir binding

1

Competition Binding

FP 6-FAM, 6 amino acid p53-derivedb peptide (Kd = 2 nM)

Trx-His6-factor Xa cleavage site-[17-125]DM2

• 1 nM FAM peptide • 10 nM DM2 • 384-well plates • Non-Langmuir binding. • Cubic equation to determine Ki

c

2

Competition Binding

TR-FRET biotin-trx-[1-83]p53; streptavidin (SA)-XL665

GST-[1-188]DM2; Eu(K)-anti-GST

• 10 nM p53 • 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

3

Phenotypic ‘GRIP’ Protein distributione

EGFP-[1-312]p53 PDE4A4-[1-124]DM2 • Internal system controls and Nutlin-3 assay positive controls.f

• 384-well plates

4

Phenotypic PCA p53-Venus[2] Venus[1]-DM2 • Transient transfection • 96-well plates

5

Direct Binding

Tm shift ‘ThermoFluor’

not applicable Glycine-[17-125]DM2 • 10 uM DM2 • 100 uM fluorescent probe, Dapoxyl

sulfonic acid • 384-well plates • FP assay used to confirm actives

6

Supplementary Table 2. Assays developed to measure the p53-DM2 interaction

Page 31: Review Inglese NatChemBiol 2007 Assays SuppTable

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).