You are on page 1of 28

FL46CH13-Gregory ARI 3 September 2013 10:50

V I E W
E
Review in Advance first posted online
R

S
on Se[te,ber 11, 2013. (Changes may
still occur before final publication
online and in print.)

C E
I N

A
D V A

Fast Pressure-Sensitive Paint


for Flow and Acoustic
Diagnostics
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

James W. Gregory,1 Hirotaka Sakaue,2 Tianshu Liu,3


and John P. Sullivan4
by University of Virginia on 10/13/13. For personal use only.

1
Department of Mechanical and Aerospace Engineering, The Ohio State University, Columbus,
Ohio 43210; email: gregory.234@osu.edu
2
Aerodynamics Research Group, Japan Aerospace Exploration Agency, Tokyo 182-8522, Japan
3
Department of Mechanical and Aeronautical Engineering, Western Michigan University,
Kalamazoo, Michigan 49008-5343
4
Department of Aeronautics and Astronautics, Purdue University, West Lafayette,
Indiana 47907-1971

Annu. Rev. Fluid Mech. 2014. 46:303–30 Keywords


The Annual Review of Fluid Mechanics is online at surface pressure measurement, porous binder, dynamic response,
fluid.annualreviews.org
high-speed imaging
This article’s doi:
10.1146/annurev-fluid-010313-141304 Abstract
Copyright  c 2014 by Annual Reviews. The development and capabilities of fast-responding pressure-sensitive paint
All rights reserved
(fast PSP) are reviewed within the context of recent applications to aerody-
namic and acoustic investigations. PSP is an optical technique for determin-
ing surface pressure distributions by measuring changes in the intensity of
emitted light, whereas fast PSP is an extension applicable to unsteady flows
and acoustics. Most fast PSP formulations are based on the development of
porous binders that allow for rapid oxygen diffusion and interaction with the
chemical sensor. This article reviews the development of porous binders, the
selection of luminophore molecules suitable for unsteady testing, dynamic
calibrations of PSP, data-acquisition methods, and noteworthy applications
for flow and acoustic diagnostics. Calibrations of the dynamic response of
fast PSP show a flat frequency response to at least 6 kHz, with some paint for-
mulations exceeding a response of 1 MHz. Various applications of fast PSP
are discussed that highlight the capabilities of the technique, and concluding
remarks highlight the need for the future development of fast PSP.

303

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

1. INTRODUCTION
Fast-responding pressure-sensitive paint (fast PSP) is an optical method for measuring time-
Pressure-sensitive resolved surface pressure distributions on a model of interest. The technique is instrumental for
paint (PSP): optical fluid dynamic, aerodynamic, and acoustic investigations owing to its high spatial resolution and
technique for high frequency response. These characteristics are particularly important for the validation of
measuring the surface
computational fluid dynamics simulations, in which the veracity of turbulence models and their
pressure based on the
oxygen modulation of results must be compared to high-resolution, time-resolved experimental data. High spatial res-
emitted light olution also makes it possible to determine surface pressures on locations of the model that are
Lifetime method: difficult to instrument with conventional pressure taps, such as sharp trailing edges, small models,
measurement of the corners with small radii, and rotating surfaces. The frequency response of fast PSP enables tem-
surface pressure by the porally and spatially resolved measurements of pressure features at rates of the order of kilohertz,
ratio of two gated yielding important insight into flow and acoustic problems that cannot be resolved by discrete
images acquired within
pressure sensors.
the exponential decay
of paint luminescence The PSP technique was first developed in the late 1980s and early 1990s (Peterson & Fitzgerald
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

1980, Kavandi et al. 1990, Mosharov et al. 1997) and has come into widespread use for wind-tunnel
Radiometric method:
pressure measurement testing (Liu et al. 1997, Bell et al. 2001, Liu & Sullivan 2005). It was only recently, however, that
by University of Virginia on 10/13/13. For personal use only.

by the ratio of wind-off fast PSP has been developed and implemented on a large scale (Gregory et al. 2008). The aim of
and wind-on images to this review is to provide an introduction to the fundamental principles of fast PSP and a concise
cancel effects of paint representation of the state of the art in fast PSP development and application. A brief outline of this
inhomogeneities and
review is as follows. After an introduction to PSP in general, along with an overview of the critical
light nonuniformity
features of fast PSP, in Section 1, Section 2 focuses on the dynamic response mechanisms of the
various constituents of fast PSP formulations. Based on this understanding of the critical dynamic
response features, we describe the development of paint formulations from the standpoint of the
chemical and material composition in Section 3. Section 4 describes methods for calibrating the
frequency response of PSP, with characterizations of the key paint formulations currently in use.
Section 5 presents several approaches to instrumentation and data-acquisition techniques for ob-
taining unsteady pressure data with fast PSP. Section 6 presents several illustrative applications of
fast PSP to unsteady flows and acoustic problems, a discussion that is naturally only a small vignette
of the broad range of modern applications of fast PSP. Section 7 summarizes the key characteristics
of fast PSP and provides our view of critical research topics to be addressed in the future.
The basic operating principles of PSP have been detailed by Liu et al. (1997), Bell et al. (2001),
and Liu & Sullivan (2005); as such, they are only briefly discussed here. PSP comprises an oxygen-
sensitive luminescent molecule (known as a luminophore) and some type of binder material to hold
the luminophore on the test article. The luminophore is excited to a heightened energy state by the
absorption of photons from an illumination source. In a simplified view, the energy of the excited
state can be released (relaxed to the ground state) through a number of mechanisms, including
Stokes-shifted luminescence at a longer wavelength, oxygen quenching, and thermal deactivation.
Thus, there is competition between the primary energy transfer mechanisms of luminescence
and oxygen quenching, which is the fundamental operating principle of PSP. A higher oxygen
concentration leads to increased quenching and reduced luminescence from the paint, making the
paint’s luminescent intensity a function of the partial pressure of oxygen in air.
There are two primary methods for acquiring the pressure information of PSP, based on
the photophysical response of the luminescent molecules: intensity and lifetime methods. The
intensity-based (radiometric) method employs continuous illumination and records the intensity
of luminescence from the paint. The raw intensity measured from PSP cannot be directly converted
to pressure because of the dependence of the measured intensity on the viewing angle and distance,
and the ambiguity resulting from spatially inhomogeneous illumination, the paint thickness, and

304 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

the luminophore concentration. Thus, a reference image at known pressure conditions is required,
and a ratio of the two images is formed. This intensity ratio is largely insensitive to the above error
sources, as long as the model remains in a fixed location. The intensity ratio may be related to
pressure via the well-known Stern-Volmer relationship (Stern & Volmer 1919, Kavandi et al.
1990),
Iref P
= A+ B , (1)
I Pref
where Iref and I are the intensity images at the reference (wind-off ) and test (wind-on) conditions,
A and B are temperature-dependent calibration coefficients, and P and Pref are the pressure distri-
butions corresponding to the test and reference conditions. Instrumentation commonly used for
the intensity method involves high-intensity, high-stability, steady light-emitting diode (LED)
illumination, with luminescence collected via a CCD (charge-coupled device) or CMOS (compli-
mentary metal oxide semiconductor) camera.
The second primary method for data acquisition is based on the dependence of the luminescent
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

lifetime on pressure. According to the kinetics of luminescence, the response of the luminescent
intensity I to time-varying illumination E(t) can be modeled by a first-order system,
by University of Virginia on 10/13/13. For personal use only.

dI I
= − + E(t), (2)
dt τ
where τ is the luminescent lifetime (Liu & Sullivan 2005, Lakowicz 2006, Gregory et al. 2008).
For pulsed illumination E(t) ∼ δ(t), the luminescent response is simply an exponential decay; in
other words,
I (t) = I0 exp(−t/τ ). (3)
Because of the oxygen quenching, the relation between the luminescent lifetime and air pressure
is also described by the Stern-Volmer equation
τref /τ = A + B(P/Pref ). (4)

In the lifetime method, the response to pulsed excitation is typically recorded through two
gated images at selected times during the exponential decay of luminescence. The initial gate is
kept very short, within the first few microseconds of the decay curve, when there is little variation in
intensity owing to changes in pressure. The second gate follows a short time later, which captures
most of the variation of the tail of the exponential decay that is more sensitive to pressure. The
key advantage of the lifetime method is that reference information is available during the test run.
Despite this, the lifetime method still requires a wind-off reference condition owing to spatial
variations in the PSP sensitivity (Goss et al. 2000). The lifetime method with wind-off correction
may be expressed as a ratio of ratios; in other words,
(IG2 /IG1 )ref P
= A+ B , (5)
IG2 /IG1 Pref
where IG1 and IG2 are the time-integrated intensity images for gates 1 and 2 of the decay curve,
respectively. Instrumentation commonly used for lifetime PSP involves a pulsed excitation source
such as an LED array or pulsed laser, along with a gated CCD camera such as an interline-transfer
CCD camera, which is often used for particle image velocimetry (PIV).
There are several common error sources in PSP methodology that we briefly mention for
context (see Liu et al. 2001a for a rigorous treatment of PSP uncertainty). The predominant
error sources are the inherent temperature sensitivity of PSP and model movement/deformation
between the test and reference conditions. The temperature problem has proven to be one of
the most vexing issues. This is often effectively addressed through in situ calibration of PSP via

www.annualreviews.org • Fast Pressure-Sensitive Paint 305

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

pressure taps or the use of a temperature map acquired via other image techniques [temperature-
sensitive paint (TSP) or infrared cameras] to correct for nonuniform temperature. Temperature
errors are also present for fast PSP; in fact, this error source tends to be accentuated because of the
TSP:
temperature-sensitive structure of the paint. The model movement problem can be partially addressed through image
paint registration (Bell & McLachlan 1996, Venkatakrishnan 2004), which effectively maps the wind-on
Biluminophore image onto the wind-off reference image for one-to-one pixel correspondence. However, if the
pressure-sensitive model moves or deflects through a nonuniform illumination field, or if model motion is along the
paint: axis of incident illumination, the spatial variation and magnitude of incident illumination will vary.
pressure-sensitive This can be counteracted by the lifetime method, which captures variations in the illumination
paint with two
field by the pressure-insensitive gate 1 image. The other primary correction scheme for model
emission wavelengths,
one pressure sensitive motion is to use a two-color paint (a biluminophore PSP), in which the second color is produced
and the other pressure by a pressure-insensitive reference probe. Model motion is a particularly relevant error source
insensitive for unsteady tests with fast PSP on rotating helicopter blades or aeroelastic wings (Gregory et al.
2009, Juliano et al. 2011, Sakaue et al. 2013a).
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

2. DYNAMIC RESPONSE MECHANISMS


by University of Virginia on 10/13/13. For personal use only.

The key physical parameters governing the dynamic response of the paint are the luminescent
lifetime and the rate of gas diffusion within the paint binder. Both must be carefully selected to
ensure a sufficiently high frequency response. Efforts to develop fast PSP for unsteady flows have
focused on improving the timescale of gas diffusion. Until recently, the binder characteristics have
been the limiting factor, but the luminescent lifetime is now becoming equally important (Fujii
et al. 2013, Sakaue et al. 2013b).

2.1. Luminescent Lifetime


The dynamic response of PSP is determined by the characteristic timescales in the photophysical
process of luminescence and the oxygen diffusion process through a polymer layer. As pointed
out above, the luminescent lifetime is the physical quantity to measure to determine the surface
pressure in the lifetime method. However, the luminescent lifetime of PSP imposes an ultimate
physical limit for an achievable temporal resolution in unsteady PSP measurements. The lifetime
of PSP is typically 1–50 μs in ambient conditions, depending on the luminophore, binder, and
solvent used. When the lifetime of a PSP is larger than the characteristic timescale of moving
features, distinct features with large pressure gradients would be blurred in PSP-derived pressure
fields. Thus, the PSP lifetime should generally be faster than the shortest characteristic timescale
of the unsteady pressure field.

2.2. Gas Diffusion Time Constant


The timescale of oxygen diffusion through a PSP layer determines the dynamic response of PSP
in unsteady measurements as the diffusion timescale is usually much larger than the luminescent
lifetime. In a thin homogeneous PSP layer, the oxygen concentration can be described by the
one-dimensional, unsteady diffusion equation with the zero-flux condition at a solid wall. The
solutions for the PSP response to changes in the oxygen concentration were given by Carroll
et al. (1996), Mosharov et al. (1997), Winslow et al. (2001), Liu & Sullivan (2005), and Gregory
& Sullivan (2006). These solutions give a classical square-law estimate for the diffusion timescale
through a homogeneous PSP layer (i.e., τdiff ∝ h 2 /Dm , where h is the coating thickness and Dm
is the mass diffusivity of oxygen in a polymer layer). For oxygen diffusion in a silicon polymer

306 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

binder, the measured value of Dm is Dm = 1.23–1.88 × 10−9 m2 /s over a temperature range of 10–
40◦ C (Carroll et al. 1997). For a 10-μm-thick polymer layer with Dm ∼ 10−9 m2 /s, the estimated
diffusion timescale is τdiff ∼ 0.1 s. Therefore, a conventional homogeneous polymer PSP has a
SNR: signal-to-noise
slow time response; it is not suitable for high-frequency unsteady pressure measurements. ratio
Porous
2.2.1. Thickness effects. The coating thickness affects not only the dynamic response but also pressure-sensitive
the signal-to-noise ratio (SNR) of PSP as the luminescent emission is proportional to the coating paint: class of paint
thickness. When the coating thickness is decreased to improve the dynamic response, the SNR binder with an open
is proportionally reduced. Schairer (2002) estimated the optimum thickness to achieve both a structure and high
surface area, enabling
suitably high dynamic response and SNR and found that the optimum thickness is less than 5 μm
the rapid interaction of
for Dm ∼ 10−9 m2 /s. However, for such a thin PSP layer, the luminescent emission is too weak oxygen with the
for accurate surface pressure measurements. This further illustrates the difficulty in applying a luminophore
conventional homogeneous polymer PSP to unsteady measurements. AA-PSP: anodized
aluminum pressure-
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

2.2.2. Porous binders. Compared to a conventional homogeneous PSP, a porous PSP could sensitive paint
have a diffusion timescale as short as a few microseconds owing to the much enlarged air-polymer PC-PSP:
by University of Virginia on 10/13/13. For personal use only.

interface of pores (Sakaue & Sullivan 2001, Sakaue et al. 2002a). Therefore, porous PSPs are fast polymer/ceramic
PSPs suitable for unsteady measurements. Interestingly, measurements of the response times of pressure-sensitive
paint
porous PSPs have showed that the exponent in the power-law relation τdiff ∝ h q is significantly
smaller than 2 (q < 2) (Asai et al. 2001). To understand the dynamic response of a porous PSP, Liu
et al. (2001b) derived the expressions for the effective diffusivity and diffusion timescale of a polymer
layer with numerous highly convoluted tube-like pores filled with air. The fractal dimension d fr is
introduced because the length of a highly convoluted tube is no longer proportional to the linear
length scale of the tube in the vertical direction. An asymptotic estimate for the diffusion timescale
of a very porous PSP is τdiff ∝ h 2−dfr Dm−1 n−1 −1
pore rpore , where n pore is the number of pores per area
and rpore is the mean radius of pores. The exponent in this power-law relation deviates from 2 by
the fractal dimension d fr owing to the presence of the fractal pores in a polymer layer, providing
a plausible explanation for the experimental finding of q < 2 in τdiff ∝ h q for porous PSPs by
Asai et al. (2001) and Sakaue & Sullivan (2001). More importantly, this relation indicates that the
diffusion timescale is inversely proportional to the number of pores per area and the mean radius
of pores. Kameda et al. (2004) calculated the effective diffusion coefficient for gas permeation
in straight cylindrical pores in anodized aluminum (AA-)PSP. The effective diffusion coefficient
significantly increases with the mean diameter of pores, and it could reach Deff ∼ 3–15×10−6 m2 /s
for anodized aluminum with pores of 10–100-nm diameter.
Gregory & Sullivan (2006) studied the response of PSP to step changes in pressure using
a diffusion model coupled with the nonlinear Stern-Volmer equation. They found that a PSP
with a diffusion timescale longer than the characteristic timescale of a measured step change in
pressure will exhibit different responses for a step increase or step decrease in pressure. They
also estimated the diffusion coefficients of fast FIB (fluoro-isopropyl-butyl polymer) PSP (Deff ∼
6.33 × 10−10 m2 /s) and polymer/ceramic (PC-)PSP (Deff ≥ 8.4 × 10−6 m2 /s) by fitting the model
to experimental data obtained with a fluidic oscillator.

3. PAINT DEVELOPMENT
PSP consists of a luminophore and a binder, with the luminophore emitting luminescence and
the binder holding the luminophore onto a testing object. As discussed in the previous section,
the oxygen diffusion of the binder is one of the major limiting factors for unsteady PSP mea-
surements. Therefore, the binding materials for fast PSP can be categorized into two main types:

www.annualreviews.org • Fast Pressure-Sensitive Paint 307

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

Incident light
a Luminescence
b

Oxygen molecules

Porous
surface Oxygen
quenching 100 nm
M ODE L SURF AC E
Luminophore
c d
Excitation Emission
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org
by University of Virginia on 10/13/13. For personal use only.

Ceramic Luminophore
particles molecules

Polymer 1 μm
M ODE L SURF ACE

Figure 1
Schematic illustrations and scanning electron micrographs of porous pressure-sensitive paint (PSP): (a,b) anodized aluminum PSP and
(c,d ) polymer/ceramic PSP. Panels a and b taken from Sakaue (2003), and panel d taken from Scroggin (1999).

porous binders and polymer binders with very high oxygen permeability. Three major types of
luminophores have been used as probe molecules: porphyrin, pyrene, and ruthenium complexes
(Liu & Sullivan 2005). Any of these can be a candidate molecule to create a fast PSP owing to
their relatively short lifetimes. Fast PSPs based on the binders are summarized below. In addi-
tion, because the temperature sensitivity and model movement are major sources of error in PSP
measurements, temperature-insensitive PSP and biluminophore fast PSP are briefly discussed as
well.

3.1. Porous Pressure-Sensitive Paint


PSP with porous material as a binder is schematically shown in Figure 1a. Porous materials have
a large surface area to which the luminophore is directly applied. The oxygen molecules in a
test gas directly quench luminescence without having to permeate into a binder layer such that
TLC-PSP: thin-layer the response time of porous PSPs is of the order of 1 μs. Various porous materials have been
chromatography plate investigated as PSP binders, such as a thin-layer chromatography (TLC) plate (Baron et al. 1993),
pressure-sensitive hydrothermal coating (Bacsa & Gratzel 1996), sol-gel (MacCraith et al. 1995), anodized aluminum
paint (Asai 1997, Sakaue 1999), anodized titanium, polymer/ceramic (Scroggin 1999), and porous filter
(Erausquin 1998). TLC-PSP, AA-PSP, and PC-PSP are the most widely used porous binders for
unsteady measurements.

308 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

3.1.1. Thin-layer chromatography plate pressure-sensitive paint. A TLC plate, which is


widely used for separation processes in chemical experiments, has different types of adsorbents,
such as silica gel, modified silica, alumina, cellulose, and kieselguhr, depending on the chemicals
PtTFPP: Pt(II)
to be separated. TLC as a PSP binder has the advantage that it is commercially available, and meso-tetrakis(pentaflu
the luminophore can be applied by just dipping the plate into a luminophore solution. Although orophenyl) porphine
TLC-PSP is easy to prepare, its brittle nature limits its application to simple shapes (e.g., Sakamura
et al. 2002, 2005; Gongora-Orozco et al. 2010).

3.1.2. Anodized aluminum pressure-sensitive paint. Aluminum is anodized by creating a thin


aluminum oxide layer on the surface by an electrochemical process. This layer is highly porous,
with 10- to 100-nm micropores uniformly distributed on the anodized aluminum surface (see
Figure 1b). Preparation processes of AA-PSP include the anodization of an aluminum model,
the modification of the surface by phosphoric acid, and the adsorption of the luminophore on the
anodized surface by dipping deposition. Details of the preparation process are given by Sakaue
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

(1999), Sakaue et al. (2006), and Gregory et al. (2008). The steady-state characteristics of AA-PSP,
such as the luminescent emission, pressure sensitivity, and temperature dependency, are influenced
by University of Virginia on 10/13/13. For personal use only.

by the luminophore application process (Zare-Behtash et al. 2012). The dipping deposition process
has been found to be most effective, with the key factors being the polarity index of the dipping
solvent (Sakaue 2005), the luminophore concentration in the dipping solution (Sakaue & Ishii
2010b), and the dipping duration (Sakaue & Ishii 2010a). The response timescales of AA-PSP
were characterized by measuring their luminescent lifetimes and response times to a step-like
pressure change (Sakaue et al. 2013b). AA-PSPs with organic luminophores possess a lifetime of
the order of 1 ns, whereas AA-PSPs with metal complex luminophores have a lifetime of the order
of 100 ns. Typical step response times of AA-PSP are discussed in Section 4.

3.1.3. Polymer/ceramic pressure-sensitive paint. Although the porous PSP formulations dis-
cussed above (TLC-PSP and AA-PSP) have been used for a number of unsteady flow measure-
ments, the material of the supporting matrix is a limiting factor. This led to the development of
PC-PSP, which can be sprayed onto any test article. Scroggin et al. (1999) initially developed the
polymer/ceramic formulation using a tape casting procedure, whereas Gregory et al. (2002b) fur-
ther developed the formulation for spraying. This PSP uses a binder containing ceramic particles
with a small amount of polymer (∼ 3.5% by weight) (Figure 1c,d). The luminophore can be applied
by mixing it within the binder solution, by dipping deposition on the polymer/ceramic surface, or
by overspraying a prepared binder surface. Gregory et al. (2008) have provided details on the prepa-
ration of PC-PSP; the paint is also available commercially from Innovative Scientific Solutions,
Inc. (Crafton et al. 2011). A new paint developed by Kameda et al. (2012) is similar to PC-PSP in
its use of nanoscale ceramic particles with high dispersion, but no polymer is added to their slurry
solution. The resulting sprayable PSP, with PtTFPP [Pt(II) meso-tetrakis(pentafluorophenyl)
porphine] luminophore mixed in the solution with the nanoparticles, produces a thin layer with
pressure and temperature sensitivities (−0.94% per kilopascal and −1.68% per degree Celsius)
that are similar to PC-PSP.
PC-PSP components such as the luminophore, polymer, and the ceramic particle determine the
characteristic outputs: the signal level, pressure sensitivity, temperature dependency, and response
time (Scroggin 1999). With bathophen ruthenium as a luminophore, the signal level can reach
an optimum output by controlling the polymer-to-particle ratio (polymer content) (Sakaue et al.
2011). The pressure sensitivity with bathophen ruthenium ranges from −0.21% to −0.95% per
kilopascal and is influenced by all the components of PC-PSP (Scroggin 1999, Gregory et al. 2006,
Gregory & Sullivan 2006, Sakaue et al. 2011). The temperature dependency is characterized by the

www.annualreviews.org • Fast Pressure-Sensitive Paint 309

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

selection of polymer and porous particles as well as the polymer content (Gregory & Sullivan 2006,
Sakaue et al. 2011). With bathophen ruthenium as a luminophore, the temperature sensitivity is
from −0.65% to −1.35% per degree Celsius. Bathophen ruthenium has historically been used
most often with PC-PSP because of its short lifetime; however, Scroggin (1999) and Juliano et al.
(2011) have shown that platinum porphyrin has a pressure sensitivity of approximately −0.82% per
kilopascal, much higher than that of the same polymer/ceramic binder with bathophen ruthenium
applied (−0.2% per kilopascal). Meanwhile, PtTFPP has a similar temperature sensitivity and
a luminescent lifetime that is only slightly longer. The response characteristics of PC-PSP are
detailed in Section 4.

3.2. Polymer-Based Fast Pressure-Sensitive Paint


An alternative to porous PSP formulations is the implementation of polymer binders with very high
oxygen permeability (Asai et al. 2001, 2002). Poly(TMSP) [poly(1-trimethylsilyl-1-propyne)] has
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

a gas permeability that is approximately 10 to 50 times greater than conventional polymer binders
such as silicone and polystyrene (Masuda et al. 1983, Amao et al. 2000, Nagai et al. 2001, Amao
by University of Virginia on 10/13/13. For personal use only.

et al. 2002). The oxygen permeation of poly(TMSP) is extremely high even at low temperatures
because of its amorphous structure with a large frozen volume. PSP using poly(TMSP) as a binder
can be fabricated in the same way as a conventional PSP; the luminophore and poly(TMSP) are
mixed in a solvent, and the resultant mixture is sprayed onto a testing object. The response time
of poly(TMSP) PSP is of the order of 10 ms (Asai et al. 2001).

3.3. Temperature-Insensitive Pressure-Sensitive Paint


Temperature-induced intensity changes are one of the most dominant error sources in PSP mea-
surement (Liu et al. 2001a). This vexing issue can be addressed by simultaneously measuring
pressure and temperature on a model and correcting the pressure errors using the known temper-
ature field and temperature sensitivity of the PSP. Another approach is the development of PSP
formulations that minimize the PSP’s temperature sensitivity. Both these techniques are described
as follows.

3.3.1. Intermediate range of two luminescent peaks. Some luminescent molecules such as
1-pyrenesulfonic acid exhibit an interesting property in which two peaks within the lumines-
cence spectrum have opposite temperature dependencies. A temperature compensation scheme
may be developed in which the luminescence is recorded over an intermediate wavelength range
that straddles the two peaks. Thus, the detected luminescence is temperature insensitive because
the temperature-induced changes of the two peaks may be cancelled out by careful selection of
optical band-pass filter cutoff wavelengths. Sakaue et al. (2012) demonstrated this concept by ap-
plying 1-pyrenesulfonic acid to AA-PSP. The resultant AA-PSP can be used for unsteady flow
measurements when a repeat test for acquiring the temperature information is not available.

3.3.2. Biluminophore fast pressure-sensitive paint. For the intensity-based method, one can
compensate for the temperature dependence of PSP by coating one-half of a symmetrical model
with PSP and the other half with TSP. Therefore, TSP data are used for the temperature com-
pensation. However, this method gives a pressure map of only one side of the coated surface under
the assumption that the flow is symmetric. Biluminophore PSP, in which PSP and TSP are mixed,
has been developed to solve these problems. The signals from PSP and TSP components can be
separated by using appropriate band-pass filters (Oglesby et al. 1996, Coyle & Gouterman 1999,

310 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

Hradil et al. 2002, Mitsuo et al. 2003, Zelelow et al. 2003, Goss et al. 2005, Kose et al. 2005,
Kimura et al. 2006, Basu et al. 2009, Stich et al. 2010, Iijima & Sakaue 2012, Long & Gregory
2012, Peng et al. 2013). Instead of correcting the temperature dependency, one can use a reference
luminescence to compensate for the pressure-independent luminescence (Khalil et al. 2004, Klein
et al. 2010a, Peng et al. 2013, Sakaue et al. 2013a). However, with biluminophore PSP, it is still a
challenge to separate the two luminescent signals. Most biluminophore PSPs involve overlapped
spectra of two luminescent signals that may cause fluorescence resonance energy transfer (Stich
et al. 2010), which reduces the pressure sensitivity. By using a porous material or a polymer with
high gas permeability, as discussed in Section 3.1, biluminophore PSP provides a fast time re-
sponse to the pressure change (Hyakutake et al. 2006, Klein et al. 2010a, Iijima & Sakaue 2012,
Peng et al. 2013, Sakaue et al. 2013a).

4. DYNAMIC CALIBRATION
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

4.1. Methods
by University of Virginia on 10/13/13. For personal use only.

Dynamic calibration methods have been developed to determine the response times of PSPs.
An ideal dynamic calibration device should provide a known pressure field at sufficiently high
frequency, for which the PSP response can be directly compared to a known standard. Dynamic
calibrations should provide information about the signal amplitude and phase delay as a function
of frequency, thus establishing the limits of the PSP response or enabling dynamic compensators.
The most common unsteady calibration device is the shock tube (Asai et al. 2001; Sakaue &
Sullivan 2001; Kameda et al. 2004, 2005, 2012; Pastuhoff et al. 2010). The shock wave created by
bursting a diaphragm provides a rapid step change in pressure. The magnitude of the pressure jump
is related to the speed of the propagating shock wave, which depends on the pressure differential
between the driver and driven sections. In calibration, a PSP sample is mounted flush with the
wall of the shock tube, allowing the shock wave to pass over the paint surface without disturbance.
A reference pressure transducer is mounted in the wall of the shock tube, and a laser is used for
illumination in point measurements of the PSP signal. The luminescent emission from PSP is
detected with a photomultiplier tube (PMT), and the response of PSP is recorded as the shock
passes the measurement point (Sakaue & Sullivan 2001). Alternatively, an image-intensified high-
speed CCD camera may be used with full-field illumination to image the passage of the shock wave
(Asai et al. 2001, Fujii et al. 2013). The shock tube is well suited for response time measurements
because it creates a large-amplitude pressure change quite rapidly. However, frequency response
data such as the phase delay and amplitude attenuation are not obtainable.
The solenoid valve is another simple device to generate a sudden pressure change for PSP
response time calibration (Baron et al. 1993; Carroll et al. 1996; Mosharov et al. 1997; Asai et al.
2001; Winslow et al. 2001; Mérienne et al. 2004, 2012). An enclosed test cell is constructed with
a PSP sample placed inside at a given pressure. The solenoid valve, which is connected to an
external supply of a different pressure, is quickly opened to create a step change in pressure. The
opening time of a typical solenoid valve is of the order of a few hundred microseconds. Solenoid
valves have the advantage of being commercially available, and the test cell is relatively simple to
construct.
Fluidic oscillators have been used by Gregory et al. (2002a,b, 2006) as a dynamic calibration
device. The frequency of the oscillating jet increases with the supply pressure, with typical oper-
ating ranges of the order of 1–10 kHz. The size of the oscillator also determines the oscillation
frequency, with smaller devices producing higher-frequency oscillations. The advantage of the
fluidic oscillator is that it can provide high dynamic pressures and high frequencies. If a reference

www.annualreviews.org • Fast Pressure-Sensitive Paint 311

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

transducer is placed in the flow field, then the PSP response may be directly compared to the
known pressure time history. When used for dynamic calibration purposes, the PSP sample
can be positioned such that the oscillating jet impinges on the PSP surface. A similar device is
Ru(dpp):
tris(4,7-diphenyl-1,10- a pulsating jet developed by Sakamura et al. (2002) for dynamic calibrations of PSP. Pulsating
phenanthroline) jets are created by chopping a jet of fluid with a chopper wheel as a high-dynamic pressure
ruthenium(II) source testing the dynamic response of the PSP sample. The pulsating frequency is limited by
bis(hexafluorophos- the dimensions, the number of cut slots, and the rotational speed of the chopper wheel. Current
phate)
calibration facilities are limited to pulsing frequencies of 1.5 kHz. However, this type of device
can provide phase and amplitude data in the frequency domain.
An acoustic standing wave tube is another device recently used for dynamic calibrations, as
the amplitude attenuation and phase delay may be directly measured as a function of frequency
(McGraw et al. 2003, Sakaue 2003, Virgin et al. 2005, Sugimoto et al. 2012, McMullen et al. 2013,
Peng et al. 2013). Pressure fluctuations introduced by a mechanical shaker (Klein et al. 2010a)
are a similar concept that may be grouped into this category. PSP is typically applied to an end
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

cap and excited by high-intensity continuous illumination, with the paint luminescence collected
from the entire end cap into a PMT. Because the pressure field along the length of the resonance
by University of Virginia on 10/13/13. For personal use only.

tube is one dimensional, the pressure field on the end cap will be uniform and vary with time. A
reference pressure transducer is also mounted in the end cap to provide a baseline for comparison.
Careful selection of the length and diameter of the resonance tube enables one to set the tube
resonance frequency and its harmonics to cover the appropriate range of interest for PSP.

4.2. Dynamic Calibration Results


The response time of AA-PSPs has been characterized by subjecting them to a step change in
pressure generated in a shock tube, as well as with the acoustic resonance tube. For AA-PSP
with a thickness of 10 ± 1 μm, it was observed that the 90% rise times to a step change in
pressure ranged from 30 to 50 μs, which were within the tolerance range caused by the thickness
uncertainty in fabricating the anodized aluminum layer (Sakaue et al. 2002a). As discussed in
Section 2.2.2, the thickness of AA-PSP is a dominant factor in the control of the response time
(Kameda et al. 2004, Sugimoto et al. 2012, Fujii et al. 2013). Sugimoto et al. (2012) found that
the luminophore used for AA-PSP—TCPP [tetrakis(4-carboxyphenyl)porphyrin] versus Ru(dpp)
[tris(4,7-diphenyl-1,10-phenanthroline)ruthenium(II) bis(hexafluorophosphate)]—does have an
impact on the dynamic response but that temperature changes have little impact on the AA-PSP
frequency response. Other dynamic calibrations of AA-PSP with a shock tube have measured step
responses with a time constant as short as 350 ns, implying a frequency response of well over
1 MHz (Fujii et al. 2013). These dynamic calibrations have shown that AA-PSP is the fastest PSP
currently available.
The response time of PC-PSP has also been characterized in shock tubes and acoustic resonance
tubes. The predominant factor governing PC-PSP dynamic response is the polymer content; for
a polymer content range of 2.6% to 90%, the response time varies from 10 μs to 10 s (Scroggin
1999, Gregory & Sullivan 2006, Gregory et al. 2006, Sakaue et al. 2011, Sugimoto et al. 2012,
Peng et al. 2013). Thus, most practical implementations of PC-PSP employ as little polymer as
possible (approximately 3% by weight) while maintaining a mechanically robust binder. Figure 2
shows a typical result from a dynamic calibration performed on PC-PSP with PtTFPP in acoustic
resonance tubes at two different laboratories. The −3-dB roll-off point for this paint formulation
is approximately 6 kHz (Sugimoto et al. 2012, McMullen et al. 2013). Kameda et al. (2012) used
a shock tube to calibrate their silica nanoparticle PSP and found step response times of less than
100 μs. In an in-depth study of the frequency response characteristics of PC-PSP, Sugimoto et al.

312 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

1 0
a b
0 −10

−1
−20
Gain (dB)

Phase (°)
−2
−30
−3
−40
−4
Sugimoto et al. (2012) −50
−5 McMullen et al. (2013)

−60
10 2 10 3 10 4 10 2 10 3 104
Frequency (Hz) Frequency (Hz)
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

Figure 2
Dynamic calibration of polymer/ceramic pressure-sensitive paint with a PtTFPP luminophore, acquired in an acoustic resonance tube:
by University of Virginia on 10/13/13. For personal use only.

(a) magnitude and (b) phase Bode plots showing behavior following a first-order system with a −3-dB roll-off point at 6 kHz. Data
replotted from Sugimoto et al. (2012) and McMullen et al. (2013). The fit to each data set is a combination of a half-order system and a
first-order system (Kameda 2012, McMullen et al. 2013).

(2012) found that the thickness of the polymer/ceramic formulation (over a range of 20 to 300 μm)
did not impact the response characteristics. This interesting result suggests that the luminophore
is deposited primarily only on the upper surface or that the paint is optically thick such that
only the upper surface luminescence is actually imaged. These observations merit further study
(particularly modeling) to better understand the response mechanisms of PC-PSP. Sugimoto
et al. (2012) also suggested that temperature has a small impact on the response characteristics and
that the luminescent lifetime of Ru(dpp) and PtTFPP predominantly drives the response char-
acteristics. However, Kameda (2012) showed that an accurate model of the PSP response should
account for both the luminescence and the diffusion timescales if the two are of the same order of
magnitude. McMullen et al. (2013) experimentally verified that a model based on a combination
of luminescence and diffusion provides the best fit to available experimental data (see Figure 2).
There have also been indications that the step-response characteristics of PSP depend on the di-
rection and magnitude of the pressure jump. Winslow et al. (2001) and Gregory & Sullivan (2006)
showed through modeling that PSP will respond faster to a decrease in pressure, and McMullen
et al. (2013) provided initial experimental data from a shock tube that verify this behavior.

5. INSTRUMENTATION, DATA ACQUISITION, AND DATA ANALYSIS


Given a fast PSP, there are several methods and instrumentation setups for acquiring pressure
data. The specific data-acquisition method used for a particular test will depend on the available
instrumentation and the test needs. The following discussion treats some of the features and
tradeoffs of the various techniques: point measurement, phase averaging, high-speed imaging,
single-shot lifetime, and motion capturing. Several of these techniques (phase averaging, high-
speed imaging, and single-shot lifetime) are discussed and compared in detail by Fang et al. (2012).

5.1. Point Measurement


A small region of PSP may be excited by a laser spot and interrogated with a PMT for point
measurements. Laser illumination provides very high incident excitation light, which can greatly

www.annualreviews.org • Fast Pressure-Sensitive Paint 313

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

improve the SNR. This feature is critical in test scenarios such as acoustics, for which the pressure
fluctuations to be measured are very small (∼1 × 102 Pa) relative to the mean pressure (∼1 ×
105 Pa). The disadvantage of this setup is that it is limited to a point measurement; field pressures
may be acquired by traversing the laser spot across the surface, although simultaneous measure-
ments in more than one location would require multiple laser/PMT setups. Nakakita (2010, 2011)
used the traversing point-measurement technique on the trailing edge of an airfoil to determine
the spectral content of noise emission.

5.2. Phase Averaging


The phase-averaging technique exploits flow periodicity to improve the SNR of a measurement.
The method is predicated on the need for a consistent, periodic signal that is characteristic
of the flow feature of interest. Often this signal can be produced by a microphone or piezo-
electric pressure transducer positioned in a strategic location on the body surface. The signal can
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

then be filtered to isolate the fundamental frequency, and phase locking can be performed on that
signal. A variable delay can be set from a reference point in the signal such that PSP images are
by University of Virginia on 10/13/13. For personal use only.

acquired over a range of phase positions to reconstruct the pressure time history. This method is
one of the first techniques used for unsteady PSP data acquisition (Gregory 2002; Gregory et al.
2002a, 2006, 2007; Sakaue et al. 2002a; Yorita et al. 2010; Asai & Yorita 2011; Disotell & Gregory
2011b; Fang et al. 2011; Pastuhoff et al. 2011; Singh et al. 2011) as it works well with traditional
CCD cameras that do not have fast frame rates. The key advantage is that the SNR may be greatly
increased by the accumulation of light on the CCD sensor until full-well intensity is reached and
by the averaging of a number of images to reduce noise. In principle, the number of cycles to be
averaged is unlimited, although this is practically limited by the expense of wind-tunnel opera-
tion and the possibility of other long-timescale error sources such as temperature changes and
photodegradation contaminating the results. The key disadvantage of this technique is the need
for a stable, periodic, single-frequency signal on which to phase lock the acquisition. In many
situations, it is not possible to produce a reference signal, or multiple frequencies can be present
in the flow. For example, in their investigation of a hemispherical dome, Fang et al. (2012) showed
that a priori knowledge of the flow-field structure would be needed to accurately reconstruct the
pressure time history via phase-averaging PSP techniques.

5.3. High-Speed Imaging


Over the past decade, the availability, cost, and quality of high-speed cameras have dramatically
improved, making them much more suitable for fast PSP measurements. The noise level of CMOS
cameras is improving to the point where the SNR is on par with that of some CCD cameras. Im-
agers capable of frame rates exceeding 10 kHz at full-frame resolution (order of megapixel and
higher) are now commonly available. High-intensity excitation is required for reasonable SNR to
be achieved; fortunately, LED illumination technology is also rapidly evolving and improving. The
key advantage of the high-speed imaging method is that real-time pressure transients—even single
events—may be captured, with no stipulation of flow periodicity. Furthermore, the availability
of a large spatial array of simultaneously acquired, time-resolved pressure enables implementa-
tion of insightful data-processing schemes such as cross-correlations (Crafton et al. 2011), spatial
maps of spectral content (Nakakita 2007), reduced-order models of surface pressure distribu-
tions (Gordeyev et al. 2013, Pastuhoff et al. 2013), and acoustic beam forming. The principal
disadvantage of high-speed imaging is the low SNR relative to the other techniques. Examples
of recent work with high-speed imaging have been given, for example, by Kameda et al. (2008),

314 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

Nakakita et al. (2008, 2012), Nakakita & Arizono (2009), Klein et al. (2010b), Crafton et al. (2011),
Nakakita (2011), Bitter et al. (2012), Fang et al. (2012), Hayashi et al. (2012), Flaherty et al. (2013),
Mérienne et al. (2013), and Hird et al. (2013).
Single-shot lifetime
method: method for
acquiring all the
5.4. Single-Shot Lifetime pressure information
from the exponential
The single-shot lifetime method of data acquisition was initially developed primarily for rotorcraft decay curve from a
applications but is applicable in scenarios in which instantaneous snapshots of the pressure field single pulse of
are needed (Gregory et al. 2009, Juliano et al. 2011). The method uses a high-energy illumination excitation light
source such as a pulsed laser to excite the PSP and provide high SNRs. A further advantage of laser Motion-capturing
excitation is that it is collimated and can be conveyed over long working distances for use in large- method: method
using a biluminophore
scale wind-tunnel facilities. Because the spatial structure of the illumination field emanating from
pressure-sensitive
a pulsed laser varies from shot to shot (due to variations in the mode structure and laser speckle), paint to correct for
this excitation source must be used in the lifetime mode in which the reference information comes errors linked to model
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

from the same excitation pulse. As shown by Goss et al. (2000, 2005), a wind-off reference image is motion through a
still required because of spatial variations in the PSP sensitivity. The unique aspect of the single- nonuniform light field
by University of Virginia on 10/13/13. For personal use only.

shot method is that all the pressure information is measured within a single pulse of excitation
light. Paint formulations with high reflectivity such as PC-PSP, coupled with the high-energy
excitation, enable data acquisition with sufficiently high SNRs. The single-shot method then
provides time-resolved pressure information at a rate equivalent to the repetition rate of pulsed
PIV lasers. The single-shot method has been routinely used at rates of 10 Hz with conventional
PIV instrumentation (cameras and laser), and it is quite possible to use a high-repetition-rate laser
with a high-speed camera to increase the data rate. An advantage of the single-shot technique
is that no image averaging is required, leading to the generation of statistics of the pressure
field. Furthermore, single-shot lifetime PSP provides reference information that can counteract
the effects of model movement. The primary disadvantage is that current commonly available
instrumentation limits the data rate to approximately 10 Hz, which is well below the fundamental
frequency of most flow and acoustic fields of interest. Thus, phase-correlated reconstructions
must still be made, even though no phase averaging is involved. Recent implementations of the
single-shot lifetime technique include the work of Disotell & Gregory (2011a), Juliano et al. (2011,
2012), Disotell et al. (2012, 2013), Fang et al. (2012), Watkins et al. (2012), Wong et al. (2012),
and Gregory et al. (2013).

5.5. Motion Capturing


Conventional, intensity-based PSP methods rely on a wind-off reference image to counteract
the effects of nonuniform illumination and to extract pressure information (Kavandi et al. 1990,
MacLachlan et al. 1993). However, this method is susceptible to significant errors if the intensity
distribution on the model changes as a result of model movement through the illumination field.
A motion-capturing method, namely imaging a two-color PSP-coated surface with a high-speed
color camera, acquires both pressure and reference information simultaneously (Sakaue et al.
2013a). As illustrated in Figure 3, this PSP has a reference luminescence and a pressure-sensitive
luminescence at two spectral peaks with sufficient separation to match the color ranges of a color
camera. By simply performing a ratio of the two color images from each time-resolved mea-
surement, one can cancel the pressure-independent image that is dependent on the camera-PSP
distance and illumination nonuniformity. By using the porous materials discussed in Section 3,
the response time of the PSP can be improved and can be applied for unsteady measurements.

www.annualreviews.org • Fast Pressure-Sensitive Paint 315

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

a b
Signal image: Reference image: Signal Reference Green Red
pressure pressure Reference/ image: image:
dependent independent signal (red) (green) Reference: Low pressure
pressure
t1 t1 t1 independent I High pressure
Signal:
pressure
t2 dependent
t2 t2 Fast frame-rate color CCD camera λ
(Phantom V12.1)

tn PC
tn tn

Long pass filter Illumination


(~440 nm) (340 ± 50 nm)
to exclude illumination
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

Figure 3
Schematic illustrating the motion-capturing pressure-sensitive paint method. (a) A ratio formed by the green and red channels can
by University of Virginia on 10/13/13. For personal use only.

cancel the error due to model movement. (b) The two color channels correspond to the green and red filters of a color CCD
(charge-coupled device) camera. Figure adapted from Sakaue et al. (2013a).

6. APPLICATIONS
To illustrate the capabilities of fast PSP applied to flow and acoustic problems, we present several
recent applications here. These applications represent a small subset of the recent proliferation
of studies using fast PSP. These particular examples highlight the primary advantages of fast PSP
in challenging test environments: fast frequency response, high spatial resolution, sensitivity to
small pressure changes, acquisition of unsteady pressure information on rotating surfaces, and
correction for severe temperature changes.
The first vignette is an application of AA-PSP to the measurement of a shock propagating over
a circular cylinder, demonstrating the fastest PSP applied to date. Fujii et al. (2013) tailored the
formation of AA-PSP by using phosphoric acid and controlling the anodization temperature and
time to influence the pore development in the anodization process. They found that shorter pores
with larger diameters have a smaller time constant, as indicated in Section 2.2.2, and were able to
reduce the time constant to as low as 0.35 μs. Two AA-PSP samples were prepared with 1-pyrene
butyric acid as the luminophore (10-ns lifetime): Sample 1 had a pore depth of 2.2 μm, a pore
diameter of 168 μm, and a time constant of 0.36 μs, whereas sample 2 had a pore depth of 10.7 μm,
a pore diameter of 20 μm, and a time constant of 10.3 μs. Figure 4 shows PSP results of the
shock propagating over the cylinder, with the paint applied to the side wall of a shock tube. The
PSP resolves the key characteristics of the shock/cylinder interaction: the formation of a curved
reflected shock from the upstream surface of the cylinder, the curvature of the propagating shock
on the downstream side of the cylinder, and the formation of the triple point downstream. The
PSP sample with shorter and wider pores (sample 1) clearly exhibits better response characteristics
by more faithfully representing these flow features.
A second demonstration of PSP involves the application of the single-shot lifetime method
with PC-PSP to the acquisition of the time-resolved pressure distribution on a helicopter blade
in forward flight. Owing to the sinusoidal variation in relative velocity experienced by a blade in
forward flight and the unsteady pitch oscillation of the blade, there are substantial unsteady pres-
sure changes present with a timescale of the order of the blade rotation frequency. Fast PSP data
acquisition is complicated by the flapping motion the rotating blade experiences, which is inherent

316 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

a Sample
le 1 b c d

5 mm Sample 2
t = 0 μs t = 10 μs t = 20 μs t = 30 μs

e f g
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org
by University of Virginia on 10/13/13. For personal use only.

t = 40 μs t = 50 μs t = 60 μs
Pressure (kPa)

20 30 40 50 60 70 80 90 100 110 120

Figure 4
Time sequence of anodized aluminum pressure-sensitive paint measurements of shock propagation over a circular cylinder (moving
from left to right). The upper half of the images shows sample 1 (2.2-μm depth, 168-μm diameter, and 0.36-μs time constant), and the
lower half of the images shows sample 2 (10.7-μm depth, 20-μm diameter, and 10.3-μs time constant). Figure adapted with permission
from Fujii et al. (2013), courtesy of Prof. Keisuke Asai.

to the operation of the helicopter. This flapping motion requires some type of reference infor-
mation to correct the illumination field—either the lifetime technique or the motion-capturing
(biluminophore) technique. Disotell et al. (2012, 2013) demonstrated the use of the single-shot
lifetime method on a small-scale helicopter in forward flight in a 3 × 5 low-speed wind tunnel.
With the blade rotating at 82 Hz, they were able to capture transient pressure information at
various blade azimuth positions (Figure 5). A series of single-shot images (no image averaging)
was acquired at each blade azimuth such that pressure fluctuation statistics could be computed for
each blade. Two key points are evident in Figure 5. There is a substantial pressure change on
the blade from the advancing to retreating blade side, which is successfully resolved by the fast
PSP, and the root-mean-squared data indicate much higher pressure fluctuations from cycle to
cycle near the tip of the retreating blade, indicating that the pressure time history is aperiodic.
One key challenge with the single-shot lifetime technique results from the architecture of the
cameras, which have a long, open-ended second gate in the two-gate acquisition. This leads to
blurring of the blade edges owing to the length of the luminescence decay. Disotell et al. (2012,
2013) employed motion-deblurring schemes developed by Juliano et al. (2012) and Gregory et al.
(2013) to deconvolve the blurred image with a point-spread function based on the known rotation
speed and luminescent lifetime. Other successful applications of fast PSP to rotorcraft include
the work of Miyamoto et al. (2010), Watkins et al. (2012), and Wong et al. (2012). Particularly

www.annualreviews.org • Fast Pressure-Sensitive Paint 317

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

0 a

p/p∞
0.5 1.00

1.0 ψ =95°
0.95
0 b
x/cmax

0.5 0.90
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

1.0 ψ =270°
prms /<p>
0.006
by University of Virginia on 10/13/13. For personal use only.

0 c
0.004

0.5

0.002

ψ =270°
1.0 μ= 0.15
0
1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2
y/R

Figure 5
Pressure distributions on a helicopter blade in forward flight with an advance ratio of 0.15 and a rotation speed of 82 Hz (the leading
edge is at the top; counterclockwise blade rotation): (a) advancing blade pressure, (b) retreating blade pressure, and (c) root-mean-
squared pressure on the retreating blade. Data were acquired with the single-shot lifetime technique, allowing for instantaneous
pressure fields to be measured and phase-correlated statistics to be computed. Figure adapted from Disotell et al. (2013).

noteworthy is the work of Wong et al. and Watkins et al. in implementing single-shot lifetime
fast PSP methods on a rotor blade in a large-scale wind-tunnel environment.
Fast PSP has also been applied to acoustics work by McGraw et al. (2003), Sakaue (2003),
Gregory et al. (2005), Virgin et al. (2005), and Disotell & Gregory (2011a). Early work by Gregory
et al. (2005) involved phase-averaged data acquisition of the pressure fluctuations within a two-
dimensional resonance cavity driven at 1.3 kHz. Their work demonstrated the capability of PSP to
resolve unsteady pressure changes as low as several hundred pascals, albeit with substantial spatial
filtering. With improved instrumentation in follow-on work, Disotell & Gregory (2011a) were
able to acquire transient pressure information with the single-shot lifetime technique (Figure 6).
This data set required much less aggressive spatial filtering and avoided the necessity of phase
averaging, while still comparing well with the analytical solution. Instantaneous pressure images
such as that in Figure 6d showed that the pressure field was nonrepeatable, with transient pressure
waves emanating from the upper-right corner of the resonance cavity at which the speaker was
mounted.
Fast PSP has also been applied to various low-speed flow applications, which exhibit many of the
same measurement challenges as acoustic applications. The low dynamic pressure associated with

318 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

Δp (kPa)

–3 –2 –1 0 1 2 3
1.0
a b
0.8

0.6
y/Ly
0.4

0.2

0
1.0
c d
0.8
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

0.6
y/Ly
0.4
by University of Virginia on 10/13/13. For personal use only.

0.2

0
0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0
x/Lx x/Lx
Δp (kPa) Δp (kPa)

–3 –2 –1 0 1 2 3 –1 –0.5 0 0.5 1

Figure 6
Polymer/ceramic pressure-sensitive paint with the single-shot lifetime technique applied to an acoustic
cavity resonating at 1.3 kHz and a sound pressure level of 154.1 dB, with no image averaging and light spatial
filtering. Images shown in panels a–c are at phase positions of 0◦ , 90◦ , and 180◦ within the oscillation cycle;
the image in panel d shows nonrepeatable higher-frequency structures that correspond to the spatial
wavelength of the acoustic signal emanating from the speaker in the upper-right corner. Figure adapted from
Disotell & Gregory (2011a). A data animation corresponding to this figure is provided in Supplemental
Video 1 (follow the Supplemental Material link from the Annual Reviews home page at http://www.
annualreviews.org).

low-speed investigations leads to very small pressure changes on the model surface that are difficult
to resolve. Flow over cylinders with circular or square cross sections has been a common subject
of study with high-speed imaging (Nakakita 2007, Pastuhoff et al. 2013) and phase-averaging
techniques (McGraw et al. 2006, Yorita et al. 2010, Asai & Yorita 2011), with the impressive ability
to resolve small-scale pressure fluctuations and the associated spatial structure of the von Kármán
instability. In other work, Nakakita (2010, 2011) studied the sound-generation mechanisms for
the trailing-edge noise of a NACA 0012 airfoil using high-speed imaging and laser-scanning PSP
systems. Bitter et al. (2012) and Flaherty et al. (2013) used high-speed imaging to study the small
pressure changes associated with the base flow of a rocket configuration and the pressure dynamics
of cavity flows, respectively.
Figure 7 shows an example of a very-high-frequency application of PSP to a microscale flow.
Here, a sample of AA-PSP was oriented parallel to the exit of a microfluidic oscillator, which
produces an oscillating jet because of the internal fluid dynamics of the device (Tomac & Gregory
2013). The entire flow field of the oscillator measured only approximately 2 mm × 2 mm, with
rich spatial and temporal gradients that could not be resolved with traditional techniques. Several

www.annualreviews.org • Fast Pressure-Sensitive Paint 319

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

0
a b c
500
y (μm)

1,000

1,500
0 500 1,000 1,500 2,000 0 500 1,000 1,500 2,000 0 500 1,000 1,500 2,000
x (μm)

Figure 7
Anodized aluminum pressure-sensitive paint (AA-PSP) measurements of a microscale fluidic oscillator flow oscillating at 9.4 kHz, at
delay increments of 20 μs from panels a to c. Note that the jet width is 325 μm, and the length scale of the flow field is approximately
2 mm × 2 mm, highlighting the spatial resolution characteristics of pressure-sensitive paint. Figure reprinted from Gregory et al.
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

(2007). A data animation corresponding to this figure is provided in Supplemental Video 2.


by University of Virginia on 10/13/13. For personal use only.

phase-averaged instances of the 9.4-kHz oscillations of the jet illustrate the bistable position of
the jet issuing from the fluidic oscillator.
An important use of temperature-insensitive AA-PSP is in short-duration, high-enthalpy wind
tunnels in which the pressure and temperature change rapidly in time. The temperature-insensitive
AA-PSP discussed in Section 3.3.1 was applied to a 30◦ compression corner model in the Mach
7.1 Hypersonic and High Enthalpy Wind Tunnel at the University of Tokyo, Kashiwa Campus
(Morita et al. 2011). 1-Pyrenesulfonic acid was used as a luminophore to provide two luminescent
peaks that are opposite in temperature dependency. The cancellation wavelength was 430 nm
at the hypersonic flow conditions of 100◦ C and 3 kPa. The surface temperature measurement
by AA-TSP showed that the front edge of the model was heated the most; its temperature rose
up to 130◦ C within the 9-s measurement time. Despite this strong temperature gradient, the
temperature-insensitive AA-PSP successfully extracted the pressure information (Figure 8). Sim-
ilar applications of AA-PSP to aerodynamic studies in short-duration hypersonic facilities have
been widespread. Relevant investigations include those of Nakakita et al. (2000), Nakakita & Asai
(2002), Sakaue et al. (2002b), Ishiguro et al. (2007), and Yang et al. (2012a,b,c).
An example of a large-scale wind-tunnel application of fast PSP is that of Mérienne et al. (2013),
who used AA-PSP to investigate transonic buffeting effects on a civil aircraft. After acquiring
images with a high-speed camera at rates up to 1 kHz and performing image processing on a
graphics processing unit, the time-resolved pressure fields provided a rich data set for investigating
the complex transonic flow in buffet conditions. Statistics such as the root-mean-squared pressure,
maps of the power spectral density, and spatial cross-correlations provided insight into the complex
shock/boundary layer interactions present in the buffet regime. For example, the maps of the power
spectral density at 108 Hz in Figure 9 show how the shock strength, location, and structure change
with the angle of attack. Other relevant studies of transonic flutter include the work of Nakakita
& Arizono (2009) and Steimle et al. (2012). The advanced data-processing techniques used by
Mérienne et al. (2013) have similarly been employed by Crafton et al. (2011), Bitter et al. (2012),
Flaherty et al. (2013), and Hird et al. (2013).
Unsteady PSP measurements using AA-PSP were conducted on a simplified rocket fairing
model in a transonic unsteady flow (Nakakita et al. 2012). A time series of pressure coefficient (Cp )
images was reduced from the unsteady PSP data measured by a high-speed camera using an in

320 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

Pressure
(kPa)
12

10

0
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

Figure 8
by University of Virginia on 10/13/13. For personal use only.

Temperature-insensitive anodized aluminum pressure-sensitive paint (AA-PSP) pressure data on a 30◦


compression corner model at Mach 7.1 (top), compared to a side view schlieren image of the shock structure
(bottom). Despite strong time-varying temperature gradients in this short-duration flow, the AA-PSP was
able to faithfully reproduce the pressure distribution. Figure reproduced from Morita et al. (2011). A data
animation corresponding to this figure is provided in Supplemental Video 3.

situ calibration based on pressure transducer data. Figure 10 shows a pressure coefficient field on
a simplified rocket fairing model and the pressure coefficients at three locations as a function of
time. An interesting feature of this work is that the PSP data reveal the complex three-dimensional
structure of the shock/boundary layer interaction, as well as the precise location of the shock as a
function of time with much better spatial resolution than the taps can provide.
Klein et al. (2010b) used PC-PSP to study the surface pressures resulting from pitch oscillations
at rates up to 30 Hz of a two-dimensional airfoil (NLR 7301) in transonic flow. They applied a
phase-locked PSP technique for measurements of the periodic phenomena. With excitation by a
single strong flash of less than 200-μs duration, the PSP generated sufficient luminescent emission
for acquisition by the CCD cameras with no need for image averaging. Figure 11 illustrates the
pressure coefficient distribution on the suction surface of the NLR 7301 airfoil in pitch oscillation
at Mach 0.72 and an angle of attack of 1.12◦ ± 0.6◦ . The PSP data show the effects of the tunnel
boundary layers on the pressure distribution, inducing three-dimensional curvature to the shocks.
The two shocks present on the surface move substantially in the chordwise direction and exhibit
complex interactions.

7. CONCLUSION AND OUTLOOK


Fast PSP has recently emerged as a viable method for resolving unsteady surface pressures in a wide
array of aerodynamic and acoustic investigations. Developments in both paint formulations and
instrumentation have enabled this significant extension in PSP capability. Paint development has
included thin polymer layers, porous binders, and hybrid formulations with the goal of improving
the gas diffusion coefficient. New formulations routinely exhibit a flat frequency response of the
order of 6 kHz, with some specific formulations (e.g., AA-PSP) showing a response as high as

www.annualreviews.org • Fast Pressure-Sensitive Paint 321

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

U∞
Temperature
sensors

Kulite
sensors

mm
130
Pa2/Hz
0 30,000
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org
by University of Virginia on 10/13/13. For personal use only.

0.54° 0.69° 0.87° 1.06° 1.23° 1.46°

Figure 9
Map of the power spectral density at 108 Hz from anodized aluminum pressure-sensitive paint (AA-PSP) data on a transonic civil
aircraft wing at M = 0.875 at varying angles of attack from 0.54◦ to 1.46◦ . The diagram of the PSP location on the wing shows that the
flow is from left to right and that the PSP sample covers only a portion of the wing at which the unsteady shock structure resides. The
PSP shows two regions of elevated pressure fluctuations owing to unsteady shock/shock and shock/boundary layer interactions, and an
aft movement of the primary shock structure as the angle of attack increases. Figure adapted from Mérienne et al. (2013), with
permission of the American Institute of Aeronautics and Astronautics. A data animation corresponding to this figure is provided in
Supplemental Video 4.

1 MHz. Advances in instrumentation include low-noise, high-speed cameras, and high-intensity


stable excitation sources. These developments have enabled data acquisition with real-time imag-
ing and advanced signal processing such as spatial spectral analysis and cross-correlations. The
past decade has also seen a progressive movement of fast PSP from benchtop investigations to
implementation in large-scale wind tunnels.
The outlook for fast PSP depends on further advancements in paint formulations and
instrumentation. Key remaining challenges include more robust compensation methods for
the simultaneous correction of temperature and illumination errors (e.g., Fischer et al. 2012,

322 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

0
PSP-tap4
Kulite-4
–0.5

Cp –1.0

–1.5

–2.0
0
PSP-tap3
Kulite-3
–0.5
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

Cp –1.0
by University of Virginia on 10/13/13. For personal use only.

–1.5

–2.0
0
PSP-tap2
Flow Kulite-2
–0.5

Cp –1.0

–1.5

–2.0
0 0.02 0.04 0.06 0.08 0.10
Time (s)

Figure 10
Map of pressure coefficients on a rocket fairing model in unsteady transonic flow. The left image shows an instant of the PSP data
demonstrating the shock/boundary layer interaction, and the right plots show the pressure coefficients at three locations as a function of
time. Figure adapted with permission from Nakakita et al. (2012). A data animation corresponding to this figure is provided in
Supplemental Video 5.

Meier et al. 2013), and paint development schemes to reduce the temperature sensitivity
(Egami et al. 2013). Paint development for enhanced sensitivity is also a critical goal as fast PSP
is applied to low-level acoustic pressure fluctuations. Whereas the frequency response of PSP is
becoming well established, more detailed modeling should be performed to better understand the
response mechanisms. As more applications involve the use of high-speed imaging for fast PSP,
there will be further implementations of advanced data-processing schemes to extract meaning
from these rich data sets. Finally, there may be interesting extensions of fast PSP, such as the recent
creation of fast-responding, pressure-sensitive microspheres that can be used for the simultaneous
measurement of the field pressure and velocity (Kimura et al. 2008).

www.annualreviews.org • Fast Pressure-Sensitive Paint 323

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

TWG side wall Cp


0.5
U∞

– 0.5

–1.5
α = 0.588 α = 1.828 α = 1.564

Figure 11
Pressure coefficient distribution on the suction surface of a two-dimensional wing-profile model (NLR 7301
model) in pitch oscillation obtained by using polymer/ceramic pressure-sensitive paint at Mach 0.72. Figure
adapted from Klein et al. (2010b), with permission from Springer-Verlag Heidelberg. A data animation
corresponding to this figure is provided in Supplemental Video 6.
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org
by University of Virginia on 10/13/13. For personal use only.

SUMMARY POINTS
1. Fast PSP is capable of measuring unsteady surface pressure distributions with high spatial
and temporal resolution.
2. Most fast PSP formulations employ porous binders, which enhance the diffusivity of the
paint.
3. The dynamic response of typical porous PSP is of the order of several kilohertz: PC-PSP
has a flat response up to 6 kHz, whereas an optimized formulation of AA-PSP has a step
response of less than 350 ns.
4. The pressure sensitivity of fast PSP is high enough such that it is suitable for some acoustic
problems such as studies of cavity resonance and airfoil trailing-edge noise generation.
5. High-speed imaging has enabled real-time measurements of pressure fields, with ad-
vanced statistical and image processing tools yielding greater insight into the flow
dynamics.
6. Methods such as single-shot lifetime, motion capturing, and biluminophore PSP have
been developed to address the most common error sources for fast PSP (temperature
and illumination errors).
7. Fast PSP has been used to study a wide array of problems, including shock dynamics,
microscale oscillating jets, rotating helicopter blades, acoustic resonance, hypersonic
flows, transonic flutter, and shock/boundary layer interactions.

FUTURE ISSUES
1. The temperature sensitivity of fast PSP is high, leading to large temperature-induced
errors. Efforts must be made to reduce the temperature sensitivity through paint
development.

324 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

2. For successful implementation on a wider range of acoustics problems, the pressure


sensitivity must be improved.
3. Effective compensation schemes for the simultaneous correction of temperature and
illumination errors are needed.
4. Further detailed modeling of the dynamic response mechanisms of porous binders will
enable the development of faster paint formulations.
5. Innovative methods for data analysis of the rich data sets resulting from fast PSP will
lead to greater insight into fluid and acoustic investigations.
6. The implementation of fast PSP must be pushed toward larger-scale wind tunnels and
flight test environments.
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

DISCLOSURE STATEMENT
J.W.G. has collaborated with Innovative Scientific Solutions, Inc., on contracts through the Small
by University of Virginia on 10/13/13. For personal use only.

Business Innovative Research (SBIR) program from NASA, the US Air Force, and the US Army.

ACKNOWLEDGMENTS
The authors gratefully acknowledge the assistance of Kevin J. Disotell and Di Peng in preparing
this review. The authors also wish to thank Keisuke Asai, Christian Klein, Marie-Claire Mérienne,
Kazuyuki Nakakita, and Aaron Scroggin for generously sharing their data, figures, and animations.
This review is dedicated to the memory of Sergey Fonov (1951–2012), whose brilliant contribu-
tions to the field of pressure-sensitive paint will long be remembered.

LITERATURE CITED
Amao Y, Asai K, Okura I, Shinohara H, Nishide H. 2000. Platinum porphyrin embedded in poly(1-
trimethylsilyl-1-propyne) film as an optical sensor for trace analysis of oxygen. Analyst 125:1911–14
Amao Y, Okura I, Shinohara H, Nishide H. 2002. An optical sensing material for trace analysis of oxygen:
metalloporphyrin dispersed in poly(1-trimethylsilyl-1-propyne) film. Polym. J. 34:411–17
Asai K. 1997. Luminescent coating with an extremely high oxygen sensitivity at low temperatures. Japanese
Patent No. 3101671
Asai K, Amao Y, Iijima Y, Okura I, Nishide H. 2002. Novel pressure-sensitive paint for cryogenic and unsteady
wind-tunnel testing. J. Thermophys. Heat Transf. 16:109–15
Asai K, Nakakita K, Kameda M, Teduka K. 2001. Recent topics in fast-responding pressure-sensitive paint
technology at National Aerospace Laboratory. Proc. 19th Int. Congr. Instrum. Aerosp. Simul. Facil., pp. 25–
36. Piscataway, NJ: IEEE
Asai K, Yorita D. 2011. Unsteady PSP measurement in low-speed flow: overview of recent advancement at Tohoku
University. Presented at AIAA Aerosp. Sci. Meet., 49th, Orlando, FL, AIAA Pap. 2011-0847
Bacsa RR, Gratzel M. 1996. Rutile formation in hydrothermally crystallized nanosized titania. J. Am. Ceram.
Soc. 79:2185–88
Baron AE, Danielson JDS, Gouterman M, Wan JR, Callis JB, McLachlan B. 1993. Submillisecond response
times of oxygen-quenched luminescent coatings. Rev. Sci. Instrum. 64:3394–402
Basu BJ, Vasantharajan N, Raju C. 2009. A novel pyrene-based binary pressure sensitive paint with low
temperature coefficient and improved stability. Sens. Actuators B 138:283–88
Bell JH, McLachlan BG. 1996. Image registration for pressure sensitive paints. Exp. Fluids 22:78–86
Bell JH, Shairer ET, Hand LA, Mehta RD. 2001. Surface pressure measurements using luminescent coatings.
Annu. Rev. Fluid Mech. 33:155–206

www.annualreviews.org • Fast Pressure-Sensitive Paint 325

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

Bitter M, Hara T, Hain R, Yorita D, Asai K, Kähler C. 2012. Characterization of pressure dynamics in
an axisymmetric separating/reattaching flow using fast-responding pressure-sensitive paint. Exp. Fluids
53:1737–49
Carroll BF, Abbitt JD, Lukas EW, Morris MJ. 1996. Step response of pressure-sensitive paints. AIAA J.
34:521–26
Carroll BF, Winslow AN, Setzer F. 1997. Mass diffusivity of pressure sensitive paints via system identification.
Presented at AIAA Aerosp. Sci. Meet., 35th, Reno, NV, AIAA Pap. 1997-0771
Coyle L, Gouterman M. 1999. Correcting lifetime measurements for temperature. Sens. Actuators B 61:92–99
Crafton J, Forlines A, Palluconi S, Hsu K-Y, Carter C, Gruber M. 2011. Investigation of transverse jet
injections in a supersonic crossflow using fast responding pressure-sensitive paint. Presented at AIAA Appl.
Aerodyn. Conf., 29th, Honolulu, AIAA Pap. 2011-3522
Disotell KJ, Gregory JW. 2011a. Measurement of transient acoustic fields using a single-shot pressure-sensitive
paint system. Rev. Sci. Instrum. 82:075112
Disotell KJ, Gregory JW. 2011b. Unsteady surface signature of a pulsed vortex generator jet. J. Vis. 14:121–27
Disotell KJ, Juliano TJ, Peng D, Gregory JW, Crafton J, Komerath NM. 2012. Unsteady pressure-sensitive
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

paint measurements on an articulated model helicopter in forward flight. Presented at AIAA Aerodyn. Meas.
Technol. Ground Test. Flight Test. Conf., 28th, New Orleans, LA, AIAA Pap. 2012-2757
Disotell KJ, Juliano TJ, Peng D, Gregory JW, Crafton J, Komerath NM. 2013. Single-shot temperature- and
by University of Virginia on 10/13/13. For personal use only.

pressure-sensitive paint measurements on an unsteady helicopter blade. Exp. Fluids. Manuscript submitted
Egami Y, Fujii K, Takagi T, Matsuda Y, Yamaguchi H, Niimi T. 2013. Reduction of temperature effects in
pressure-sensitive paint measurements. AIAA J. In press
Erausquin RG. 1998. Cryogenic temperature- and pressure-sensitive fluorescent paints. MS thesis. Sch. Aeronaut.
Astronaut., Purdue Univ., West Lafayette, IN
Fang S, Disotell KJ, Long SR, Gregory JW, Semmelmayer FC, Guyton RW. 2011. Application of fast-
responding pressure-sensitive paint to a hemispherical dome in unsteady transonic flow. Exp. Fluids
50:1495–505
Fang S, Long SR, Disotell KJ, Gregory JW, Semmelmayer FC, Guyton RW. 2012. Comparison of unsteady
pressure-sensitive paint measurement techniques. AIAA J. 50:109–22
Fischer LH, Karakus C, Meier RJ, Risch N, Wolfbeis OS, et al. 2012. Referenced dual pressure- and
temperature-sensitive paint for digital color camera read out. Chem. Eur. J. 18:15706–13
Flaherty W, Reedy TM, Elliott GS, Austin JM, Schmit RF, Crafton J. 2013. Investigation of cavity flow using
fast-response pressure sensitive paint. Presented at AIAA Aerosp. Sci. Meet., 51st, Grapevine, TX, AIAA
Pap. 2013-0678
Fujii S, Numata D, Nagai H, Asai K. 2013. Development of ultrafast-response anodized-aluminum pressure-sensitive
paints. Presented at AIAA Aerosp. Sci. Meet., 51st, Grapevine, TX, AIAA Pap. 2013-0485
Gongora-Orozco N, Zare-Behtash H, Kontis K. 2010. Global unsteady pressure-sensitive paint measurements
of a moving shock wave using thin-layer chromatography. Measurement 43:152–55
Gordeyev S, De Lucca N, Jumper E, Hird K, Juliano TJ, et al. 2013. The comparison of unsteady pressure field
over flat- and conformal-window turrets using pressure sensitive paint. Presented at AIAA Fluid Dyn. Conf.,
43rd, San Diego, CA
Goss L, Jones G, Crafton J, Fonov S. 2005. Temperature compensation for temporal (lifetime) pressure sensitive
paint measurements. Presented at AIAA Aerosp. Sci. Meet. Exhib., 43rd, Reno, NV, AIAA Pap. 2005-1027
Goss LG, Trump DD, Sarka B, Lydick LN, Baker WM. 2000. Multi-dimensional time-resolved pressure-sensitive
paint techniques: a numerical and experimental comparison. Presented at AIAA Aerosp. Sci. Meet., 38th, Reno,
NV, AIAA Pap. 2000-0832
Gregory JW. 2002. Unsteady pressure measurements in a turbocharger compressor using porous pressure-sensitive
paint. MS thesis. Sch. Aeronaut. Astronaut., Purdue Univ., West Lafayette, IN
Gregory JW, Asai K, Kameda M, Liu T, Sullivan JP. 2008. A review of pressure-sensitive paint for high-speed
and unsteady aerodynamics. Proc. IMechE G J. Aerosp. Eng. 222:249–90
Gregory JW, Juliano TJ, Disotell KJ, Peng D, Crafton J, Komerath NM. 2013. Deconvolution-based algorithms
for deblurring PSP images of rotating surfaces. Presented at AIAA Aerosp. Sci. Meet., 51st, Grapevine, TX,
AIAA Pap. 2013-0484

326 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

Gregory JW, Kumar P, Peng D, Fonov S, Crafton J, Liu T. 2009. Integrated optical measurement techniques for
investigations of fluid-structure interactions. Presented at AIAA Fluid Dyn. Conf., 39th, San Antonio, TX,
AIAA Pap. 2009-4044
Gregory JW, Sakaue H, Sullivan JP. 2002a. Fluidic oscillator as a dynamic calibration tool. Presented at Aerodyn.
Meas. Technol. Ground Test. Conf., 22nd, St. Louis, MO, AIAA Pap. 2002-2701
Gregory JW, Sakaue H, Sullivan JP. 2002b. Unsteady pressure measurements in turbomachinery using porous
pressure sensitive paint. Presented at AIAA Aerosp. Sci. Meet., 40th, Reno, NV, AIAA Pap. 2002-0084
Gregory JW, Sullivan JP. 2006. Effect of quenching kinetics on unsteady response of pressure-sensitive paint.
AIAA J. 43:634–45
Gregory JW, Sullivan JP, Raghu S. 2005. Visualization of jet mixing in a fluidic oscillator. J. Vis. 8:169–76
Gregory JW, Sullivan JP, Raman G, Raghu S. 2007. Characterization of the microfluidic oscillator. AIAA J.
45:568–76
Gregory JW, Sullivan JP, Wanis SS, Komerath NM. 2006. Pressure-sensitive paint as a distributed optical
microphone array. J. Acoust. Soc. Am. 119:251–61
Hayashi T, Ishikawa H, Sakaue H. 2012. Development of polymer-ceramic pressure-sensitive paint and its
application to supersonic flow field. 28th Int. Symp. Shock Waves, Vol. 1, ed. K Kontis, pp. 607–13. Berlin:
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

Springer
Hird K, Juliano TJ, Gregory JW, Gordeyev S, De Lucca N, et al. 2013. Study of unsteady surface pressure on a
by University of Virginia on 10/13/13. For personal use only.

turret via pressure-sensitive paint. Presented at AIAA Fluid Dyn. Conf., 43rd, San Diego, CA
Hradil J, Davis C, Mongey C, McDonagh C, MacCraith BD. 2002. Temperature-corrected pressure-
sensitive-paint measurements using a single camera and a dual-lifetime approach. Meas. Sci. Technol.
13:1552–57
Hyakutake T, Taguchi H, Sakaue H, Nishide H. 2006. Oxygen visualization membranes composed of polypyridyl-
propyne and porphyrins. Presented at Int. Symp. Polym. Chem., Dalian, China, P65
Iijima Y, Sakaue H. 2012. Platinum porphyrin and luminescent polymer for pressure- and temperature-sensing
probes. Sens. Actuators A 184:128–33
Ishiguro Y, Ohmi S, Nagai H, Asai K, Nakakita K. 2007. Visualization of hypersonic compression corner flows using
temperature- and pressure-sensitive paints. Presented at AIAA Aerosp. Sci. Meet., 45th, Reno, NV, AIAA
Pap. 2007-0118
Juliano TJ, Disotell KJ, Gregory JW, Crafton J. 2012. Motion-deblurred, fast-response pressure-sensitive
paint on a rotor in forward flight. Meas. Sci. Technol. 23:045303
Juliano TJ, Kumar P, Peng D, Gregory JW, Crafton S, Fonov S. 2011. Single-shot, lifetime-based pressure-
sensitive paint for rotating blades. Meas. Sci. Technol. 22:085403
Kameda M. 2012. Effect of luminescence lifetime on the frequency response of fast-response pressure-sensitive
paints. Trans. Jpn. Soc. Mech. Eng. B 78:1942–50 (In Japanese)
Kameda M, Seki H, Makoshi T, Amao Y, Nakakita K. 2008. Unsteady measurement of a transonic delta wing flow
by a novel PSP. Presented at AIAA Appl. Aerodyn. Conf., 26th, Honolulu, AIAA Pap. 2008-6418
Kameda M, Seki H, Makoshi T, Amao Y, Nakakita K. 2012. A fast-response pressure sensor based on a
dye-adsorbed silica nanoparticle film. Sens. Actuators B 171–172:343–49
Kameda M, Tabei T, Nakakita K, Sakaue H, Asai K. 2005. Image measurements of unsteady pressure fluctu-
ation by a pressure-sensitive coating on porous anodized aluminum. Meas. Sci. Technol. 16:2517–74
Kameda M, Tezuka N, Hangai T, Asai K, Nakakita K, Amao Y. 2004. Adsorptive pressure-sensitive coatings
on porous anodized aluminum. Meas. Sci. Technol. 15:489–500
Kavandi JL, Callis JB, Gouterman MP, Khalil G, Wright D, et al. 1990. Luminescent barometry in wind
tunnels. Rev. Sci. Instrum. 61:3340–47
Khalil G, Costin C, Crafton J, Jones G, Grenoble S, et al. 2004. Dual luminophor pressure sensitive paint: I.
Ratio of reference to sensor giving a small temperature dependency. Sens. Actuators B 97:13–21
Kimura F, Khalil G, Zettsu N, Xia Y, Callis J, et al. 2006. Dual luminophore polystyrene microspheres for
pressure-sensitive luminescent imaging. Meas. Sci. Technol. 17:1254–60
Kimura F, Rodriguez M, McCann J, Carlson B, Dabiri D, et al. 2008. Development and characterization of
fast responding pressure sensitive microspheres. Rev. Sci. Instrum. 79:074102
Klein C, Sachs WE, Henne U, Borbye J. 2010a. Determination of transfer function of pressure-sensitive paint.
Presented at AIAA Aerosp. Sci. Meet., 48th, Orlando, FL, AIAA Pap. 2010-0309

www.annualreviews.org • Fast Pressure-Sensitive Paint 327

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

Klein C, Sachs WE, Henne U, Egami Y, Mai H, et al. 2010b. Application of pressure-sensitive paint for
determination of dynamic surface pressures on a 30 Hz oscillating 2D profile in transonic flow. In New
Results in Numerical and Experimental Fluid Mechanics VII, ed. A Dillmann, G Heller, M Klaas, HP Kreplin,
pp. 323–30. Notes Numer. Fluid Mech. Multidiscip. Des. Vol. 112. New York: Springer
Kose ME, Carroll BF, Schanze KS. 2005. Preparation and spectroscopic properties of multiluminophore
luminescent oxygen and temperature sensor films. Langmuir 21:9121–29
Lakowicz JR. 2006. Principles of Fluorescence Spectroscopy. New York: Springer. 3rd ed.
Liu T, Campbell B, Bruns S, Sullivan JP. 1997. Temperature- and pressure-sensitive luminescent paints in
aerodynamics. Appl. Mech. Rev. 50:227–46
Liu T, Guille M, Sullivan JP. 2001a. Accuracy of pressure-sensitive paint. AIAA J. 39:103–12
Liu T, Sullivan JP. 2005. Pressure and Temperature Sensitive Paints. Berlin: Springer
Liu T, Teduka N, Kameda M, Asai K. 2001b. Diffusion timescale of porous pressure-sensitive paint. AIAA J.
39:2400–2
Long SR, Gregory JW. 2012. A temperature corrected dual-luminophore pressure-sensitive paint system for unsteady
pressure measurements. Presented at AIAA Aerosp. Sci. Meet., 50th, Nashville, AIAA Pap. 2012-0131
MacCraith BD, McDonagh CM, O’Keeffe G, McEvoy AK, Butler T, Sheridan FR. 1995. Sol-gel coatings for
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

optical chemical sensors and biosensors. Sens. Actuators B 29:51–57


MacLachlan BG, Kavandi JL, Callis JB, Gouterman MP, Green E, Khalil G. 1993. Surface pressure field
by University of Virginia on 10/13/13. For personal use only.

mapping using luminescent coatings. Exp. Fluids 14:33–41


Masuda T, Isobe E, Higashimura T, Takada K. 1983. Poly[1-(trimethylsilyl)-1-propyne]: a new high polymer
synthesized with transition-metal catalysts and characterized by extremely high gas permeability. J. Am.
Chem. Soc. 105:7473–74
McGraw CM, Bell JH, Khalil G, Callis JB. 2006. Dynamic surface pressure measurements on a square cylinder
with pressure sensitive paint. Exp. Fluids 40:203–11
McGraw CM, Shroff H, Khalil G, Callis JB. 2003. The phosphorescence microphone: a device for testing
oxygen sensors and films. Rev. Sci. Instrum. 74:5260–66
McMullen RM, Huynh DP, Gregory JW, Crafton J. 2013. Dynamic calibrations for fast-response porous poly-
mer/ceramic pressure-sensitive paint. Presented at AIAA Ground Test. Conf., San Diego, CA
Meier RJ, Fischer LH, Wolfbeis OS, Schäferling M. 2013. Referenced luminescent sensing and imaging with
digital color cameras: a comparative study. Sens. Actuators B 177:500–6
Mérienne MC, Coponet D, Luyssen JM. 2012. Transient pressure-sensitive paint investigation in a nozzle.
AIAA J. 50:1453–61
Mérienne MC, Le Sant Y, Ancelle J, Soulevant D. 2004. Unsteady pressure measurement instrumentation
using anodized-aluminum PSP applied in a transonic wind tunnel. Meas. Sci. Technol. 15:2349–60
Mérienne MC, Le Sant Y, Lebrun F, Deleglise B, Sonnet D. 2013. Transonic buffeting investigation using unsteady
pressure-sensitive-paint in a large wind tunnel. Presented at AIAA Aerosp. Sci. Meet., 51st, Grapevine, TX,
AIAA Pap. 2013-1136
Mitsuo K, Asai K, Hayasaka M, Kameda M. 2003. Temperature correction of PSP measurement using dual-
luminophor coating. J. Vis. 6:213–23
Miyamoto K, Miyazaki T, Sakaue H. 2010. Development of a motion-cancelled PSP system and its application to a
helicopter blade. Presented at AIAA Aerodyn. Meas. Technol. Ground Test. Conf., 27th, Chicago, AIAA
Pap. 2010-4798
Morita K, Suzuki K, Imamura O, Sakaue H. 2011. Temperature-cancelled anodized-aluminum pressure sensitive
paint for hypersonic wind tunnel application. Presented at AIAA Fluid Dyn. Conf. Exhib., 41st, Honolulu,
AIAA Pap. 2011-3724
Mosharov VE, Radchenko VN, Fonov SD. 1997. Luminescent Pressure Sensors in Aerodynamic Experiments.
Moscow: Cent. Aerohydrodyn. Inst. (TsAGI), CWA Int. Corp.
Nagai K, Masuda T, Nakagawa T, Freeman BD, Pinnau I. 2001. Poly[1-(trimethylsilyl)-1-propyne] and
related polymers: synthesis, properties and functions. Prog. Polym. Sci. 26:721–98
Nakakita K. 2007. Unsteady pressure distribution measurement around 2D-cylinders using pressure-sensitive paint.
Presented at AIAA Appl. Aerodyn. Conf., 25th, Miami, FL, AIAA Pap. 2007-3819
Nakakita K. 2010. Scanning unsteady PSP technique for high-frequency and small-pressure fluctuation in low-speed.
Presented at AIAA Aerodyn. Meas. Technol. Ground Test. Conf., 27th, Chicago, AIAA Pap. 2010-4920

328 Gregory et al.

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

Nakakita K. 2011. Unsteady pressure measurement on NACA0012 model using global low-speed unsteady PSP
technique. Presented at AIAA Fluid Dyn. Conf., 41st, Honolulu, AIAA Pap. 2011-3901
Nakakita K, Arizono H. 2009. Visualization of unsteady pressure behavior of transonic flutter using pressure-sensitive
paint measurement. Presented at AIAA Appl. Aerodyn. Conf., 27th, San Antonio, TX, AIAA Pap. 2009-
3847
Nakakita K, Asai K. 2002. Pressure-sensitive paint application to a wing-body model in a hypersonic shock tunnel.
Presented at AIAA Aerodyn. Meas. Technol. Ground Test. Conf., 22nd, St. Louis, MO, AIAA Pap. 2002-
2911
Nakakita K, Osawa J, Hori N, Kameda M. 2008. Unsteady pressure-sensitive paint measurement for oscillating shock
wave in supersonic nozzle. Presented at AIAA Appl. Aerodyn. Conf., 26th, Honolulu, AIAA Pap. 2008-6580
Nakakita K, Takama Y, Imagawa K, Kato H. 2012. Unsteady PSP measurement of transonic unsteady flow field
around a rocket fairing model. Presented at AIAA Aerodyn. Meas. Technol. Ground Test. Flight Test.
Conf., 28th, New Orleans, LA, AIAA Pap. 2012-2758
Nakakita K, Yamazaki T, Asai K, Teduka N, Fuji A, Kameda M. 2000. Pressure sensitive paint measurement in
a hypersonic shock tunnel. Presented at AIAA Aerodyn. Meas. Technol. Ground Test. Conf., 21st, Denver,
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

AIAA Pap. 2000-2523


Oglesby DM, Upchurch BT, Leighty BD, Simmons KA. 1996. Pressure sensitive paint with internal temper-
ature sensing luminophore. Proc. 42nd Int. Instrum. Symp. San Diego: Instrum. Soc. Am.
by University of Virginia on 10/13/13. For personal use only.

Pastuhoff M, Tillmark N, Alfredsson PH. 2010. Dynamic calibration of polymer/ceramic pressure sensitive
paint using a shock tube. Proc. 7th Int. Conf. Flow Dyn., pp. 132–33. Sendai, Jpn: Tohoku Univ.
Pastuhoff M, Tillmark N, Alfredsson PH. 2011. Wall pressure measurements in a Y-junction at pulsating
flow using polymer/ceramic pressure sensitive paint. Proc. 10th Int. Symp. Exp. Comput. Aerothermodyn.
Internal Flows, Pap. ISAIF10-155. Brussels: Vrije Univ.
Pastuhoff M, Yorita D, Asai K, Alfredsson PH. 2013. Enhancing the signal-to-noise ratio of pressure sensitive
paint data by singular value decomposition. Meas. Sci. Technol. 24:075301
Peng D, Jensen CD, Juliano TJ, Gregory JW, Crafton J, et al. 2013. Temperature-compensated fast pressure-
sensitive paint. AIAA J. In press. doi: 10.2514/1.J052318
Peterson JI, Fitzgerald RV. 1980. New technique of surface flow visualization based on oxygen quenching of
fluorescence. Rev. Sci. Instrum. 51:670–71
Sakamura Y, Matsumoto M, Suzuki T. 2005. High frame-rate imaging of surface pressure distribution using
a porous pressure-sensitive paint. Meas. Sci. Technol. 16:759–65
Sakamura Y, Suzuki T, Matsumoto M, Masuya G, Ikeda Y. 2002. Optical measurements of high-frequency
pressure fluctuations using a pressure-sensitive paint and Cassegrain optics. Meas. Sci. Technol. 13:1591–98
Sakaue H. 1999. Porous pressure sensitive paints for aerodynamic applications. MS thesis. Sch. Aeronaut. Astronaut.,
Purdue Univ., West Lafayette, IN
Sakaue H. 2003. Anodized aluminum pressure sensitive paint for unsteady aerodynamic applications. PhD diss. Sch.
Aeronaut. Astronaut., Purdue Univ., West Lafayette, IN
Sakaue H. 2005. Luminophore application method of anodized aluminum pressure sensitive paint as a fast
responding global pressure sensor. Rev. Sci. Instrum. 76:084101
Sakaue H, Gregory JW, Sullivan JP, Raghu S. 2002a. Porous pressure-sensitive paint for characterizing
unsteady flowfields. AIAA J. 40:1094–98
Sakaue H, Ishii K. 2010a. Dipping duration study for optimization of anodized-aluminum pressure-sensitive
paint. Sensors 10:9799–807
Sakaue H, Ishii K. 2010b. Optimization of anodized-aluminum pressure-sensitive paint by controlling lu-
minophore concentration. Sensors 10:6836–47
Sakaue H, Kakisako T, Ishikawa H. 2011. Characterization and optimization of polymer-ceramic pressure-
sensitive paint by controlling polymer content. Sensors 11:6967–77
Sakaue H, Kuriki T, Miyazaki T. 2012. A temperature-cancellation method of pressure-sensitive paint on
porous anodic aluminum. J. Lumin. 132:256–60
Sakaue H, Matsumura S, Schneider SP, Sullivan JP. 2002b. Anodized aluminum pressure sensitive paint for short
duration testing. Presented at AIAA Aerodyn. Meas. Technol. Ground Test. Conf., 22nd, St. Louis, MO,
AIAA Pap. 2002-2908

www.annualreviews.org • Fast Pressure-Sensitive Paint 329

Changes may still occur before final publication online and in print
FL46CH13-Gregory ARI 3 September 2013 10:50

Sakaue H, Miyamoto K, Miyazaki T. 2013a. A motion-capturing pressure-sensitive paint method. J. Appl.


Phys. 113:084901
Sakaue H, Morita K, Iijima Y, Sakamura Y. 2013b. Response time scales of anodized-aluminum pressure-
sensitive paints. Sens. Actuators A 199:74–79
Sakaue H, Sullivan JP. 2001. Time response of anodized aluminum pressure-sensitive paint. AIAA J. 39:1944–
49
Sakaue H, Tabei T, Kameda M. 2006. Hydrophobic monolayer coating on anodized aluminum pressure-
sensitive paint. Sens. Actuators B 119:504–11
Schairer ET. 2002. Optimum thickness of pressure-sensitive paint for unsteady measurements. AIAA J.
40:2312–18
Scroggin AM. 1999. Processing and optimization of luminescence based pressure and temperature sensors for aerody-
namic applications. MS thesis. Sch. Mater. Sci. Eng., Purdue Univ., West Lafayette, IN
Scroggin AM, Slamovich EB, Crafton JW, Lachendo N, Sullivan JP. 1999. Porous polymer/ceramic com-
posites for luminescent-based temperature and pressure measurement. Mater. Res. Soc. Proc. 560:347–52
Singh M, Naughton JW, Yamashita T, Nagai H, Asai K. 2011. Surface pressure and flow field behind an
Annu. Rev. Fluid Mech. 2014.46. Downloaded from www.annualreviews.org

oscillating fence submerged in turbulent boundary layer. Exp. Fluids 50:701–14


Steimle PC, Karhoff DK, Schröder W. 2012. Unsteady transonic flow over a transport-type swept wing. AIAA
J. 50:399–415
by University of Virginia on 10/13/13. For personal use only.

Stern O, Volmer M. 1919. Uber die abklingungszeit der fluoreszenz. Phys. Z. 20:183–88
Stich MIJ, Fischer LH, Wolfbeis OS. 2010. Multiple fluorescent chemical sensing and imaging. Chem. Soc.
Rev. 39:3102–14
Sugimoto T, Kitashima S, Numata D, Nagai H, Asai K. 2012. Characterization of frequency response of pressure-
sensitive paints. Presented at AIAA Aerosp. Sci. Meet., 50th, Nashville, AIAA Pap. 2012-1185
Tomac MN, Gregory JW. 2013. Jet interactions in a feedback-free fluidic oscillator at low flow rate. Presented at
AIAA Fluid Dyn. Conf., 43rd, San Diego, CA
Venkatakrishnan L. 2004. Comparative study of different pressure-sensitive-paint image registration tech-
niques. AIAA J. 42:2311–19
Virgin CA, Carroll BF, Cattafesta LN, Schanze KS, Kose ME. 2005. Pressure sensitive paint for acoustic
pressure fluctuations. Proc. 2005 ASME Int. Mech. Eng. Congr. Expo., pp. 297–307, Pap. IMECE2005-
81831. New York: ASME
Watkins AN, Leighty BD, Lipford WE, Wong OD, Goodman KZ, et al. 2012. Deployment of a pressure
sensitive paint system for measuring global surface pressures on rotorcraft blades in simulated forward flight.
Presented at AIAA Aerodyn. Meas. Technol. Ground Test. Flight Test. Conf., 28th, New Orleans, LA,
AIAA Pap. 2012-2756
Winslow NA, Carroll BF, Kurdila AJ. 2001. Model development and analysis of the dynamics of pressure-
sensitive paints. AIAA J. 39:660–66
Wong OD, Watkins AN, Goodman KZ, Crafton JW, Forlines A, et al. 2012. Blade tip pressure mea-
surements using pressure sensitive paint. Proc. 68th Am. Helicopter Soc. Annu. Forum Technol. Disp., Pap.
AHS2012-000233. Alexandria: AHS Int.
Yang L, Erdem E, Zare-Behtash H, Kontis K. 2012a. Pressure-sensitive paint on a truncated cone in hypersonic
flow at incidences. Int. J. Heat Fluid Flow 37:9–21
Yang L, Zare-Behtash H, Erdem E, Kontis K. 2012b. Application of AA-PSP to hypersonic flows: the double
ramp model. Sens. Actuators B 161:100–7
Yang L, Zare-Behtash H, Erdem E, Kontis K. 2012c. Investigation of the double ramp in hypersonic flow
using luminescent measurement systems. Exp. Therm. Fluid Sci. 40:50–56
Yorita D, Nagai H, Asai K, Narumi T. 2010. Unsteady PSP technique for measuring naturally-disturbed periodic
phenomena. Presented at AIAA Aerosp. Sci. Meet., 48th, Orlando, FL, AIAA Pap. 2010-0307
Zare-Behtash H, Yang L, Gongora-Orozco N, Kontis K, Jones A. 2012. Anodized aluminium pressure
sensitive paint: effect of paint application technique. Measurement 45:1902–5
Zelelow B, Khali GE, Phelan G, Carlson B, Gouterman M, et al. 2003. Dual luminophor pressure sensitive
paint: II. Lifetime based measurement of pressure and temperature. Sens. Actuators B 96:304–14

330 Gregory et al.

Changes may still occur before final publication online and in print

You might also like