You are on page 1of 9

2046 JOURNAL OF ATMOSPHERIC AND OCEANIC TECHNOLOGY VOLUME 25

A Blackbody Design for SI-Traceable Radiometry for Earth Observation


P. JONATHAN GERO, JOHN A. DYKEMA, AND JAMES G. ANDERSON
School of Engineering and Applied Sciences, Harvard University, Cambridge, Massachusetts

(Manuscript received 10 December 2007, in final form 22 April 2008)

ABSTRACT

Spaceborne measurements pinned to international standards are needed to monitor the earth’s climate,
quantify human influence thereon, and test forecasts of future climate change. The International System of
Units (SI, from the French for Système International d’Unités) provides ideal measurement standards for
radiometry as they can be realized anywhere, at any time in the future. The challenge is to credibly prove
on-orbit accuracy at a claimed level against these international standards. The most accurate measurements
of thermal infrared spectra are achieved with blackbody-based calibration. Thus, SI-traceability is obtained
through the kelvin scale, making thermometry the foundation for on-orbit SI-traceable spectral infrared
measurements. Thermodynamic phase transitions are well established as reproducible temperature stan-
dards and form the basis of the international practical temperature scale (International Temperature Scale
of 1990, ITS-90). Appropriate phase transitions are known in the temperature range relevant to thermal
infrared earth observation (190–330 K) that can be packaged such that they are chemically stable over the
lifetime of a space mission, providing robust and traceable temperature calibrations. A prototype blackbody
is presented that is compact, highly emissive, thermally stable and homogeneous, and incorporates a small
gallium melting point cell. Precision thermal control of the blackbody allows the phase transition to be
identified to within 5 mK. Based on these results, the viability of end-to-end thermometric calibration of
both single-temperature and variable-temperature blackbodies on orbit by employing multiple-phase-
change cells was demonstrated.

1. Introduction the distribution of temperature, water vapor, and


clouds, measurements are required to continuously
Obtaining accurate measurements of infrared radi- cover at least the 200–2000 cm⫺1 spectral range at a
ance from space can contribute greatly to our under- resolution of better than or equal to 1 cm⫺1 (Ohring
standing of climate change and the earth system. The
2008). In this spectral window, global observations over
long-wave forcing of the climate, the climate’s re-
all seasons will encounter radiance temperatures from
sponse, and the long-wave feedbacks involved in that
190 to 330 K. Various studies (Anderson et al. 2004;
response bear characteristic signatures in a time series
Ohring 2008; National Research Council 2007) have
of thermal infrared spectra. The long-wave water vapor
indicated that signal detection above natural variability
feedback, cloud feedback, and temperature change are
for decadal climate signatures requires a threshold un-
uniquely discernable. For this reason, high spectral
certainty of 0.05–0.1 K (1␴), or better, in radiance tem-
resolution thermal infrared time series constitute an ef-
perature. This is equivalent to a relative combined un-
fective benchmark of global climate change. Further-
certainty of 1.5 ⫻ 10⫺3 (at 250 K at 750 cm⫺1) in radi-
more, such a time series should provide powerful con-
ance (1␴ uncertainty values are used throughout this
straints for climate models by improving the represen-
paper). Realizing such low uncertainty is routine for
tation of feedbacks (Leroy et al. 2008).
national meteorology institutes in a laboratory setting,
To capture the spectral features containing informa-
but proving on-orbit uncertainty at this level has not yet
tion about greenhouse gas forcing and the response of
been accomplished by space-based infrared sounders.
Attaining a high-accuracy measurement from a satellite
instrument that is defensible and credible by the scientific
Corresponding author address: Dr. P. Jonathan Gero, School of
Engineering and Applied Sciences, Harvard University, 12 Ox-
community will necessitate traceability to the Interna-
ford St., Cambridge, MA 02138. tional System of Units (SI, from the French for Système
E-mail: gero@huarp.harvard.edu International d’Unités; see Dykema and Anderson 2006).

DOI: 10.1175/2008JTECHA1100.1

© 2008 American Meteorological Society


Brought to you by University of Maryland, McKeldin Library | Unauthenticated | Downloaded 02/06/24 07:34 AM UTC
NOVEMBER 2008 GERO ET AL. 2047

The International System of Units is linked to fun- ted by a cavity with a finite aperture with Lambertian
damental physical properties of matter such that any reflectance is
experimenter anywhere in the world can establish the
I␯˜ ⫽ ␧␯˜ B␯˜ 共T 兲 ⫹ 共1 ⫺ ␧␯˜ 兲 B␯˜ 共T eff兲, 共2兲
bias of their observation tied to the applicable SI stan-
dard. The statement that a given measurement is SI where ␧␯˜ is the cavity spectral emissivity and T is the eff

traceable implies that it can be related to the base SI effective temperature of the radiation from the back-
units through an unbroken chain of comparisons with ground environment. The second term on the right-
accepted standards, all having stated uncertainties. In hand side of this equation is a simplification valid for
addition, certain principles must be followed to evalu- spatially isotropic and isothermal background radia-
ate the uncertainty with which data are reported. For a tion. The values of the physical constants that appear in
satellite instrument, SI-traceable calibration achieved the Planck function are known with much lower uncer-
during prelaunch calibration cannot be assumed to be tainties than is required for remote sensing applica-
valid during the operational lifetime of the instrument. tions. Thus, the dominant source of uncertainty in a
Demonstrating SI traceability on orbit is challenging well-designed blackbody is in the measurement of the
because accepted standards with which the full dynamic cavity temperature, as well as the effect of the nonunity
range of expected measurements can be calibrated are emissivity of a practical blackbody with a macroscopic
not readily available, and the component uncertainties aperture. Temperature probes can exhibit drift over
of the measurement (exclusive of the calibration stan- time periods of years, readout electronics can degrade,
dards) are not easily quantified. Demonstrating SI and cavity surface preparations can be altered though
traceability on orbit, however, is essential, since it is the oxidation in the low earth orbit environment. The re-
only rigorous path to establishing a high-accuracy time alization of the Planck function can thus become inac-
series. curate following preflight calibration. Current and
There are two common methods to realize an SI- planned operational infrared sounders have no means
traceable infrared radiance scale. One is detector based to directly measure this drift; therefore, they have no
through electrical substitution radiometry, which is means to ascertain the accuracy of their calibration tar-
based on SI electrological standards and the principle gets during the flight mission. This results in a calibra-
of equivalence between the optical watt and the elec- tion error of unknown magnitude that varies with time
trical watt (Martin et al. 1985). The other is source on orbit.
based through the use of blackbodies with SI-traceable To address this problem, we have developed a black-
thermometry whose radiance can be described by the body design that implements SI-traceable thermom-
Planck function for blackbody radiation (Fox 2000). Al- etry, throughout the lifetime of a satellite instrument,
though national meteorology institutes have built their by realizing the primary SI temperature scale on orbit.
primary radiometric scales around the detector-based The base SI unit of temperature is the kelvin, and the
method, on spaceborne infrared sounders lower uncer- basic meteorological temperature scale is the Interna-
tainties may be achieved with source-based methods. A tional Temperature Scale of 1990. ITS-90 is defined in
review of remote sensing calibration by National Insti- relation to the fixed points of pure elements, which are
tute of Standards and Technology (NIST) meteorolo- immutable physical constants (Rusby et al. 1991). Our
gists suggests that large systematic errors can occur in design incorporates a simple gallium melting point stan-
practical blackbodies (Rice and Johnson 2001). The at- dard into a calibration blackbody. The gallium melting
tainment of the low uncertainties required for climate point remains constant over time and is a defining point
measurements therefore necessitates blackbodies with of the ITS-90. It is used as a benchmark to calibrate the
an SI-traceable emission scale. onboard thermometers. This method allows accurate
The spectral radiance B␯˜ (T ) emitted by a blackbody determination of a blackbody temperature during the
cavity of uniform temperature T with an infinitesimal lifetime of the satellite instrument on orbit, directly
aperture is described by the Planck function: traceable to the base SI temperature scale. To fully
constrain the blackbody radiance, an accurate determi-
2hc2␯˜ 3 nation of the effective cavity emissivity is also neces-

冉 冊
B␯˜ 共T 兲 ⫽ , 共1兲
hc␯˜ sary. This is the subject of Gero et al. (2008; manuscript
exp ⫺1 submitted to J. Atmos. Oceanic Technol.). Here, we
kBT
describe the operation of the blackbody with an em-
where h is Planck’s constant, c is the speed of light in a bedded melting point standard for calibrating ther-
vacuum, kB is the Boltzmann constant, and ␯˜ is the mometers. In section 2 we describe the details of the
spectral index (in cm⫺1). The spectral radiance I␯˜ emit- experimental apparatus, specifically the thermal and

Brought to you by University of Maryland, McKeldin Library | Unauthenticated | Downloaded 02/06/24 07:34 AM UTC
2048 JOURNAL OF ATMOSPHERIC AND OCEANIC TECHNOLOGY VOLUME 25

FIG. 1. Cross-sectional view of the blackbody with the embedded gallium cell (shaded in
gray). The blackbody has cylindrical geometry. All component modules (except the gallium
cell) are made of aluminum 6061 and the cavity interior is coated with Aeroglaze Z306. The
interior length is 229 mm, the inner diameter is 51 mm, and the entrance aperture diameter
is 38 mm. All components are held together using compression. Thermal grease is used as
indicated to enhance conduction between the gallium cell and the blackbody. Thermistors,
labeled 1–6, are placed into the narrow cylindrical wells to monitor the temperature along the
length of the cavity. Thermistors 1, 2, and 6 are referred to as the gallium, cone, and aperture
thermistors, respectively.

optical design elements of the blackbody and the op- is used, with 13-mm-thick walls, in order to enhance the
eration of the gallium melting point standard. In section thermal conductivity and minimize temperature gradi-
3 we discuss the experimental results obtained with the ents along the cavity. The blackbody exterior is covered
blackbody. In section 4 we discuss the implications of with layers of Kapton, Nomex, and aluminized Mylar to
these results for implementing a three-point calibration insulate against radiative thermal losses to the environ-
of temperature sensors on orbit. We summarize our ment. The full blackbody module is mounted on an
conclusions in section 5. optical table with highly rigid and insulating G10 fiber-
glass stands to reduce thermal losses in the mounting
mechanism. Each blackbody module is outfitted with a
2. Experiment description Kapton thermofoil heater (Minco, Minneapolis, Min-
nesota). The power dissipation in each can be manually
The design strategy employed in this experiment was controlled to ensure temperature homogeneity across
to construct a blackbody that obeyed the Planck func- the entire cavity. The heaters are regulated using an
tion [Eq. (1)] with an uncertainty in radiance domi- analog proportional-integral-derivative controller (Wave-
nated by the uncertainty from temperature measure- length Electronics, Bozeman, Montana). A thermistor
ment. This strategy required that the errors in radiance placed in proximity to the reentrant cone provides the
temperature from cavity temperature nonuniformity closed-loop feedback for the controller. The opera-
and nonunity cavity emissivity be less than the ther- tional blackbody has a setpoint accuracy of 5 mK, tem-
mometry error. In the following sections we outline the perature stability better than 1 mK h⫺1, and cavity in-
thermal and optical designs, respectively, that were homogeneity of 10 mK along its length, depending on
used to meet these requirements. This is followed by a the difference between the blackbody setpoint and the
description of the gallium melting point standard that ambient temperature.
confers SI traceability to the blackbody temperature. The temperatures of each blackbody module and the
The performance of the blackbody–gallium cell system gallium cell are monitored using thermistors (Thermo-
was tested in a vacuum chamber simulating the low metrics, Edison, New Jersey). They are potted with
earth orbit environment. thermal grease in deep narrow cylindrical cavities 2–5
mm from the blackbody interior surface. The exact lo-
cations of the six thermistors used are shown in Fig. 1.
a. Thermal design The resistance of the thermistor is measured with a
Figure 1 depicts the cylindrical cavity geometry that Hart Scientific 1575 SuperThermometer (American
was chosen for the blackbody, with a 60° reentrant cone Fork, Utah). This apparatus applies a constant current
at the base. The aperture, the cone, and the cylinder of 10 ␮A and measures the voltage across the resistive
were designed as independent interlocking aluminum sensor, comparing it to the voltage across a well-
modules that are held together using compression. characterized internal reference resistor. The measure-
Temperature homogeneity is achieved by carefully con- ment is performed twice with the current in alternating
trolling several design elements. A large thermal mass directions, in order to eliminate offset voltages, includ-

Brought to you by University of Maryland, McKeldin Library | Unauthenticated | Downloaded 02/06/24 07:34 AM UTC
NOVEMBER 2008 GERO ET AL. 2049

TABLE 1. Component uncertainties (1␴) of the gallium cell b. Optical design


blackbody.
The aperture, the cone, and the cylinder were de-
Uncertainty signed as independent interlocking aluminum modules.
Component (mK)
The 60° reentrant cone at the base of the blackbody
Thermometry ensures that an incident ray undergoes a large number
Readout electronics 0.5
of specular reflections before exiting the cavity. The
Calibration standard 0.4
Calibration fitting 3.0 blackbody cavity may be assembled with an arbitrary
Blackbody number of cylindrical modules, thereby providing a
Homogeneity 0.1 range of aspect ratios, defined as the ratio of cavity
Emissivity 2.1 depth to aperture diameter. Higher aspect ratios lead to
higher effective emissivities and will more closely ap-
ing those arising from thermal electromotive forces proximate the ideal blackbody (Gouffé 1945; Quinn
(EMFs). Using this approach, errors from driving cur- 1967; Chandos and Chandos 1974; Usadi 2006). Based
rent imprecision, voltage offsets, amplifier and analog- on preliminary emissivity modeling, an aspect ratio of 6
to-digital converter (ADC) inaccuracies, and drift in was chosen, with a cavity length of 229 mm and an
the physical properties of electronic components are aperture diameter of 38 mm.
avoided because these all affect the voltage samples The interior surface of the cavity was sandblasted to
equally (Hart Scientific 1999). Lead resistance errors create micron-scale roughness. The surface was primed
are eliminated by using a four-wire circuit. In this ar- with Aeroglaze 9947 primer, and coated with Aero-
rangement, the sensor is driven with current from one glaze Z306 diffuse black paint (Lord Corp., Erie, Penn-
pair of wires and the resulting EMF is sensed with a sylvania). This paint has characteristically high emissiv-
second pair of wires. The signal is passed to an amplifier ity in the infrared, low outgassing rates, and is qualified
with very high input impedance that draws negligible for spaceborne applications. A thin 0.06-mm coat of
current from the sensor. As a result, no measurable paint was applied in order to keep temperature gradi-
voltage develops along the EMF sensing wires. Electri- ents across the insulating paint to a minimum (Best et
cal noise remains the chief source of measurement un- al. 2003).
certainty. The ultimate accuracy of the SuperTher- The directional–hemispherical reflectance of a wit-
mometer in measuring resistance is two parts in 105, or ness sample of the blackbody surface preparation was
0.5 mK between 25° and 50°C (Hart Scientific 1999). determined with a reflectometer. The apparatus em-
Table 1 lists the component uncertainty budget. ployed a quantum cascade laser at 8 ␮m in pulsed
The thermistors are individually calibrated in a mode. The hemispherically reflected laser light from a
highly stable and homogenous thermal bath (Hart Sci- surface sample was collected in a gold-coated integrat-
entific 7080 Calibration Bath). They are placed in a ing sphere (Labsphere, North Sutton, New Hampshire)
calibration fluid mixture composed of 50% water and and the signal amplitude was measured synchronously
50% ethylene glycol, and their resistance values are with a mercury–cadmium–telluride (MCT) detector
measured at seven different temperatures from ⫺30°C (Kolmar, Newburyport, Massachusetts), then com-
to 50°C. At each setpoint, the bath temperature is held pared to a surface sample with known reflectance in
constant for 30 min while the thermistor resistances are order to calculate the absolute reflectance. The surface
logged. The bath temperature is determined using a reflectance of the witness sample was measured to be
standard platinum resistance thermometer (SPRT; 0.031 ⫾ 0.001, with measurement precision being the
Hart Scientific 5681 SPRT) with a traceable calibration dominant source of uncertainty.
to the ITS-90 temperature scale. The resulting tempera- The effective emissivity of the blackbody cavity was
ture and resistance data are fitted to the Steinhart–Hart modeled using a Monte Carlo method (Sapritsky and
equation in order to determine the calibration coeffi- Prokhorov 1992; Sapritsky and Prokhorov 1995;
cients of each thermistor. The fit residuals to the Stein- Prokhorov 1998). The statistical ray-tracing calculation
hart–Hart equation are less than ⫾3 mK over the entire provides estimates of normal spectral effective emissiv-
range. ity for blackbody cavities of arbitrary cylindrical geom-
The experiment was conducted in a vacuum chamber etry and temperature distribution. The average normal
to simulate the thermal environment of low earth orbit. ray-viewing condition with a beam diameter of 25 mm
The chamber was evacuated to ⬍1 ⫻ 10⫺5 mb, where was employed to correspond to the size of the image of
convection is negligible and radiation is the dominant the detectors in the plane of the blackbody aperture.
mechanism of heat transfer. The simulations were run with 107 rays, leading to an

Brought to you by University of Maryland, McKeldin Library | Unauthenticated | Downloaded 02/06/24 07:34 AM UTC
2050 JOURNAL OF ATMOSPHERIC AND OCEANIC TECHNOLOGY VOLUME 25

FIG. 2. Cross-sectional view of the gallium melting point stan-


dard. The cell has cylindrical geometry and is 117 mm long and 23
mm in diameter. The temperature probe is inserted into the well
down the middle of the cell, where it is in good thermal contact
with the melting gallium. The cell contains 25 g of gallium with a
nominal purity of 99.999 99%. FIG. 3. A full gallium melting point transition in a thermal bath.
The cell, initially at room temperature, was immersed in a liquid
bath held at 29.92°C. The temperature inside the cell plateaued at
the gallium melting point for about 10 h, before equilibrating with
uncertainty of one part in 106 resulting from ray trun- the bath temperature.
cation.
Using the measured temperature distributions along
the cavity (10–18-mK gradient along the cylinder as standard is to begin with the gallium in its solid phase
measured by the thermistors) and a surface reflectance and immerse the cell in a liquid bath at a temperature
of 0.031 (as measured at 8 ␮m with the reflectometer), just above the melting point. As the gallium begins to
the modeled effective cavity emissivity was 0.999 883 ⫾ melt, the temperature inside the well reaches the melt-
0.000 003. The dominant source of error is the uncer- ing point and remains stable at that temperature while
tainty in the surface reflectance. For a cavity heated to the material is undergoing melting (shown in Fig. 3).
50°C in an effective ambient background of 27°C, this The length of the melting plateau is inversely propor-
results in a radiance temperature of 49.9976°C, a 2.4- tional to the temperature difference between the bath
mK deviation from an ideal blackbody. If the cavity temperature and the melting point. With the bath set to
temperature were isothermal, its radiance temperature 29.9°C, the plateau lasts approximately 10 h.
under these conditions would be 49.9977°C, 0.1 mK In this experimental setup, the laboratory blackbody
different compared to the observed case. These errors in good thermal contact with the cell acts as the tem-
are less than the 3.1-mK cumulative thermometry mea- perature bath. A mechanical mount was designed to
surement error; thus, the model results indicate that the place the cell adjacent to the reentrant cone in close
blackbody should exhibit excellent radiometric perfor- proximity to the primary emitting surface (shown in
mance. Fig. 1). Good thermal contact between the aluminum
and the cell was ensured by using thermal grease.
c. Gallium melting point standard
3. Experimental results
The SRM 1968 gallium melting point standard—http://
ts.nist.gov/MeasurementServices/ReferenceMaterials/ Figure 4 shows the temperature of the gallium cell
archived_certificates/1968.%20June%201977.pdf (NIST, and the blackbody during a full gallium phase transi-
Gaithersburg, Maryland)—consists of approximately tion. Here, the temperature of the blackbody was first
25 g of high-purity gallium in a chemically stable epoxy- stabilized below the gallium melt point at 29.18°C.
sealed Teflon crucible. The cylindrical cell is 117 mm Then, constant power was applied to the heaters to
tall and 23 mm in diameter, with the geometry shown in gradually raise the temperature above the melt point.
Fig. 2. It allows simple realization of the gallium melt- Around time index equal to 410 min, the temperature
ing point at 29.7646 ⫾ 0.0007°C, which is a defining of the gallium cell plateaued, while the surrounding
point of the ITS-90 temperature scale (Thornton 1977). blackbody continued to increase in temperature, albeit
The cell was designed for calibrating small temperature at a lower rate. The melt plateau lasted about 120 min,
probes up to 3.6 mm in diameter. after which the gallium began to warm up and eventu-
The standard operation of the gallium melting point ally reequilibrated with the rest of the blackbody.

Brought to you by University of Maryland, McKeldin Library | Unauthenticated | Downloaded 02/06/24 07:34 AM UTC
NOVEMBER 2008 GERO ET AL. 2051

FIG. 5. Onset of the gallium melting point with the blackbody


FIG. 4. A full melting transition of the gallium cell blackbody. operated in step mode. (top) The temperature of the blackbody
(top) The temperature of the blackbody cone (thermistor 2) and cone and the gallium cell; (bottom) the temperature difference
the gallium cell (thermistor 1) during the melt. The blackbody between the two. The temperature of the blackbody is stepped up
temperature was first stabilized below the gallium melting point; at 5–15-mK intervals, then allowed to equilibrate for an hour. The
then, constant power was applied to the blackbody heaters after temperature difference equilibrates to approximately 40 mK prior
approximately 370 min. The gallium melting point was reached to the melt, but increases once the melt is initiated. The transition
around time index 410 min, at which point the temperature inside point between the two regimes is clearly discernable at time index
the cell remained constant for the next 120 min. After the melt, 180 min.
the temperature of the cell reequilibrated with the rest of the
blackbody. The other parts of the blackbody (thermistors 3–6)
were all within 10 mK of each other under equilibrium conditions melting was initiated, the temperature difference be-
prior to the melt, and they reached a maximum gradient of 90 mK
tween the cone and the gallium thermistors increased,
between the aperture and the cone near the end of the melt pla-
teau. (bottom) The rate of change of temperature for the black- as the melting of the gallium perturbed the thermal
body cone and the gallium cell. equilibrium, acting as an additional reservoir for ther-
mal energy. This can be clearly identified in Fig. 5 at
time index 180 min. The temperature of the surround-
Under equilibrium conditions, a thermal gradient ing blackbody is 29.805°C at this point, 41 mK above
was present between the blackbody cone and the tem- the gallium melting point. The temperature inside the
perature inside the gallium cell. The gallium cell was cell stabilized at 29.762 88°C ⫾ 0.000 04 over the next
found to be 25–60 mK colder than the cone, under 23 h, while the surrounding blackbody was at 29.835°C.
various experimental conditions. This was partly due to The small offset from the true gallium melting point
the 7.5-mm layer of Teflon separating the aluminum was within the thermistor uncertainty (⫾3.1 mK; see
cavity and the melting material. Furthermore, because Table 1).
of the geometry, heat was only directly applied to one
end of the cylindrical cell, while the other end equili-
4. Discussion
brated with the surrounding (colder) thermal environ-
ment. Using this methodology, the accuracy of blackbody
The onset of the melt plateau was more clearly iden- thermometry can be determined on board a satellite
tified by raising the temperature of the blackbody in instrument, traceable to the ITS-90 temperature scale.
small 5–15-mK steps. In this method of operation the Since the gallium is hermetically sealed and its physical
initiation of the melting point was extremely distinct properties are not subject to long-term drift, such a cell
and could be best identified by looking at the tempera- can be used to test thermometer drift over the lifetime
ture difference between the cone and the gallium of a satellite instrument. To achieve this operationally,
thermistors (shown in Fig. 5). While the gallium was the fixed-point temperature must be related thermally
solid, this temperature difference approached a con- to the temperature probes embedded in the blackbody.
stant value that was nearly independent of the tempera- The main obstacle encountered in the laboratory ex-
ture (⬃40 mK). In this regime the equilibrium tempera- periments was the 25–60-mK gradient between the
ture distribution is determined by the thermal conduc- phase-change material and the thermistor near the ra-
tivity of the blackbody–gallium cell system. Once the diating surface. This gradient can be reduced by im-

Brought to you by University of Maryland, McKeldin Library | Unauthenticated | Downloaded 02/06/24 07:34 AM UTC
2052 JOURNAL OF ATMOSPHERIC AND OCEANIC TECHNOLOGY VOLUME 25

proving heat transfer between the blackbody and the


gallium, and by employing a mechanical design that
completely surrounds the cell within the heated black-
body. Any remaining gradient can be treated as a con-
stant bias that can be accounted for with a thermal
model of the blackbody. The exact onset of the phase
transition needs to be identified. This can be achieved
with either the step or the ramp method used in this
paper. With the step method, a thermometer must di-
rectly measure the temperature of the melt material, as
the onset is determined using the difference between
the temperature of the phase-change material and the
temperature measured by the thermometer to be cali-
brated. This requires good active control of the black-
body to maintain it at a steady temperature. The accu-
racy of realizing the temperature at the onset of the
phase transition is dependent on the smallest achiev-
able temperature step size. In this paper the realization
accuracy was 5–15 mK. With the ramp method, the
onset of the melting point can be identified by looking
at the signature of a perturbed rate of change of heating
in any thermometer embedded in the blackbody. This
method requires good thermal control of the black-
body, as the accuracy of identifying the melting point
improves with lower temperature ramp rates. It has
been demonstrated that melting point signatures can be FIG. 6. Phase-change fixed points within the radiance tempera-
ture range of earth observations of infrared radiance (190–330 K);
identified to 5-mK accuracy using a ramp rate of about
all values are in kelvins. Three defining points of the ITS-90 lie in
0.02 mK s⫺1 (Best et al. 2007). this range (indicated by circles): the triple point of mercury, the
Thermometers suitable for space applications, such triple point of water, and the melting point of gallium. Three
as thermistors and encapsulated platinum resistance eutectic alloys of gallium—GaIn, GaSn, and GaZn (indicated by
thermometers, can have their calibration coefficients diamonds)—also have well-defined melting points that may be
used in this application.
determined using a minimum of three temperature
points. The relation between the resistance R and the
temperature T for a thermistor is best described by the points of gallium and its eutectic alloys (shown in Fig. 6)
Steinhart–Hart equation (Steinhart and Hart 1968): would be suitable for this application (Krutikov et al.
2006). For a climate instrument with the mission to
T ⫺1 ⫽ A ⫹ B log共R兲 ⫹ C 关log共R兲兴3, 共3兲 measure SI-traceable radiance over a broad radiance
temperature range corresponding to various earth-
where A, B, and C are empirically derived calibration
observing conditions, a variable-temperature black-
coefficients. For a platinum resistance thermometer,
body is needed. In this case the fixed points of mercury,
the resistance–temperature relation is best approxi-
water, and gallium can be used to calibrate a tempera-
mated by the Callendar–van Dusen equation (Callen-
ture probe in the range spanning 234–303 K (Best et al.
dar 1887; van Dusen 1925):
2007).
R ⫽ R0关1 ⫹ AT ⫹ BT 2 ⫹ C共T ⫺ 100兲T 3兴, 共4兲 The overall temperature uncertainty of performing
such an in situ calibration of a thermistor with three
where R0 is the resistance at 0°C. In both cases the fixed-point cells was modeled and the results are shown
three coefficients A, B, and C can be calculated by in Fig. 7. The calculation was done with a model for
measuring a minimum of three distinct resistance– thermistor calibration using the Steinhart–Hart equa-
temperature pairs and performing a linear regression. tion, and a half-bridge topology for measuring the
For a spaceborne blackbody that operates near a thermistor resistance, similar to the one employed by
single temperature throughout its mission, closely Keith et al. (2001) in the Interferometer for Emission
spaced fixed points may be used to calibrate the ther- and Solar Absorption (INTESA) flight instrument.
mometer over a narrow temperature range. The fixed Component uncertainties arising from the readout elec-

Brought to you by University of Maryland, McKeldin Library | Unauthenticated | Downloaded 02/06/24 07:34 AM UTC
NOVEMBER 2008 GERO ET AL. 2053

TABLE 2. Component uncertainties (1␴) of the thermometry


readout electronics for a modeled flight blackbody design. The
uncertainty in the realization and identification of the fixed points
was varied over the values of 0, 5, 10, and 15 mK.

Temperature
Component 234 K 273 K 303 K
Thermistor self-heating 0.01 mK 0.11 mK 0.39 mK
Readout electronics 0.2 ⍀ 0.2 ⍀ 0.2 ⍀
Reference resistors 0.2 ⍀ 0.2 ⍀ 0.2 ⍀
ADC resolution (12 bit) 153.4 ⍀ 8.3 ⍀ 3.4 ⍀

the blackbody can be calibrated on orbit, during the


lifetime of the instrument, with traceability to the ITS-
90 temperature scale in the temperature range 234–303
K. Since the practically achievable pathway for the
FIG. 7. Combined temperature uncertainty of a modeled in situ traceability of infrared radiance to the SI is through
calibration of blackbody thermistors using three fixed-point cells. Planck’s law and the definition of the kelvin, this meth-
(top) The uncertainty in calibrating a 10-k⍀ thermistor using the odology allows unprecedented accuracy in the remote
fixed points of Hg, H2O, and Ga (indicated by the vertical lines of
sensing of spectral infrared radiance. This is an im-
symbols). (bottom) The uncertainty for a 30-k⍀ thermistor cali-
brated at the fixed points of GaSn, GaZn, and Ga. provement over previous methods for infrared radiance
calibration, which have appealed to engineering formu-
las or intercomparison campaigns capable only of test-
tronics, including reference resistors, thermistor self- ing system–level uncertainty. This methodology pro-
heating, and analog-to-digital converter resolution, vides an end-to-end calibration of all aspects of ther-
were accounted for in the model, and are listed in Table mometry, including thermometer coefficients and
2. The overall uncertainty in the temperature is evalu- electronics readout.
ated from an ensemble of 104 Monte Carlo simulations
of the calibration model, where the variance of the ran- Acknowledgments. The authors acknowledge the en-
dom variables was specified by the 1␴ root-sum-of- gineering support of L. Lapson, M. Greenberg, J. De-
squares measurement uncertainties. The results show musz, M. Rivero, and T. Martin. The authors would
that within the region bounded by the fixed points the also like to thank F. Best and his coinvestigators for
combined thermometric uncertainty is dominated by helpful discussions and S. Mekhontsev for suggestions
the uncertainty in temperature determination. These on the practical realization of the fixed-point concept.
results suggest that with sufficient accuracy in tempera-
ture determination, it is possible to calibrate tempera- REFERENCES
ture probes on a satellite instrument on orbit, with ad-
equately low uncertainties to meet the demands of cli- Anderson, J. G., J. A. Dykema, R. M. Goody, H. Hu, and D. B.
Kirk-Davidoff, 2004: Absolute, spectrally-resolved, thermal
mate observations. radiance: A benchmark for climate monitoring from space. J.
Quant. Spectrosc. Radiat. Transfer, 85, 367–383.
5. Conclusions Best, F. A., and Coauthors, 2003: Traceability of absolute radio-
metric calibration for the Atmospheric Emitted Radiance In-
An in situ fixed-point evaluation of the calibration terferometer (AERI). Proc. Characterization and Radiomet-
ric Calibration for Remote Sensing, Logan, UT, Space Dy-
uncertainty of thermometers was performed in a high-
namics Laboratory.
emissivity, thermally stable, compact blackbody with an ——, D. Adler, S. Ellington, D. Thielman, H. Revercomb, and J.
embedded gallium melting point cell. The design is Anderson, 2007: On-orbit absolute temperature calibration
evolvable into a lightweight flight version. Two meth- for CLARREO. Proc. CALCON Tech. Conf., Logan, UT,
ods have been investigated for realizing the phase tran- Space Dynamics Laboratory.
sition for the purposes of thermometer calibration. Us- Callendar, H. L., 1887: On the practical measurement of tempera-
ture: Experiments made at the Cavendish Laboratory, Cam-
ing these methods, thermometry accurate to 5 mK may
bridge. Philos. Trans. Roy. Soc. London, A178, 161–230.
be achieved. By extending the methodology employed Chandos, R. J., and R. E. Chandos, 1974: Radiometric properties
in this experiment to multiple fixed-point cells within a of isothermal, diffuse wall cavity sources. Appl. Opt., 13,
single blackbody, the temperature probes embedded in 2142–2152.

Brought to you by University of Maryland, McKeldin Library | Unauthenticated | Downloaded 02/06/24 07:34 AM UTC
2054 JOURNAL OF ATMOSPHERIC AND OCEANIC TECHNOLOGY VOLUME 25

Dykema, J. A., and J. G. Anderson, 2006: A methodology for Prokhorov, A. V., 1998: Monte Carlo method in optical radiom-
obtaining on-orbit SI-traceable spectral radiance measure- etry. Metrologia, 35, 465–471.
ments in the thermal infrared. Metrologia, 43, 287–293. Quinn, T. J., 1967: The calculation of the emissivity of cylindrical
Fox, N. P., 2000: Primary radiometric quantities and units. Metro- cavities giving near black-body radiation. Brit. J. Appl. Phys.,
logia, 37, 507–513. 18, 1105–1113.
Gouffé, A., 1945: Aperture corrections for artificial blackbodies Rice, J. P., and B. C. Johnson, 2001: NIST activities in support of
with consideration of multiple internal diffusions (in French). space-based radiometric remote sensing. Harnessing Light:
Rev. Opt., 24, 1–10. Optical Science and Metrology at NIST, C. Londono, Eds.,
Hart Scientific, 1999: 1575/1590 Super-Thermometer user’s guide. International Society for Optical Engineering (SPIE Pro-
Hart Scientific, 149 pp. ceedings, Vol. 4450), 108–126.
Rusby, R. L., R. P. Hudson, M. Durieux, J. F. Schooley, P. P. M.
Keith, D. W., J. A. Dykema, H. Hu, L. Lapson, and J. G. Ander-
Steur, and C. A. Swenson, 1991: Thermodynamic basis of the
son, 2001: Airborne interferometer for atmospheric emission
ITS-90. Metrologia, 28, 9–18.
and solar absorption. Appl. Opt., 40, 5463–5473.
Sapritsky, V. I., and A. V. Prokhorov, 1992: Calculation of the
Krutikov, V. N., and Coauthors, 2006: The Global Earth Obser-
effective emissivities of specular-diffuse cavities by the
vation System of Systems (GEOSS) and metrological support
Monte Carlo method. Metrologia, 29, 9–14.
for measuring radiometric properties of objects of observa-
——, and A. V. Prokhorov, 1995: Spectral effective emissivities of
tions. Metrologia, 43, S94–S97.
nonisothermal cavities calculated by the Monte Carlo
Leroy, S. S., J. G. Anderson, J. A. Dykema, and R. M. Goody, method. Appl. Opt., 34, 5645–5652.
2008: Testing climate models using thermal infrared spectra. Steinhart, J. S., and S. R. Hart, 1968: Calibration curves for
J. Climate, 21, 1863–1875. thermistors. Deep-Sea Res., 15, 497–503.
Martin, J. E., N. P. Fox, and P. J. Key, 1985: A cryogenic radiom- Thornton, D. D., 1977: The gallium melting point standard: A
eter for absolute radiometric measurements. Metrologia, 21, determination of the liquid-solid equilibrium temperature of
147–155. pure gallium on the International Practical Temperature
National Research Council, 2007: Earth Science and Applications Scale of 1968. Clinical Chem., 23, 719–724.
from Space: National Imperatives for the Next Decade and Usadi, E., 2006: Reflecting cavity blackbodies for radiometry.
Beyond. National Academies Press, 456 pp. Metrologia, 43, S1–S5.
Ohring, G., Ed., 2008: Achieving Satellite Instrument Calibration van Dusen, M. S., 1925: Platinum-resistance thermometry at low
for Climate Change. U.S. Dept. of Commerce, 144 pp. temperatures. J. Amer. Chem. Soc., 47, 326–332.

Brought to you by University of Maryland, McKeldin Library | Unauthenticated | Downloaded 02/06/24 07:34 AM UTC

You might also like