Sunteți pe pagina 1din 8

Cryogenics 55–56 (2013) 12–19

Contents lists available at SciVerse ScienceDirect

Cryogenics
journal homepage: www.elsevier.com/locate/cryogenics

Thermal conductivity of rigid foam insulations for aerospace vehicles


M. Barrios ⇑, S.W. Van Sciver
National High Magnetic Field Laboratory, Tallahassee, FL 32310, USA
Mechanical Engineering Department, FAMU/FSU College of Engineering, Tallahassee, FL 32310, USA

a r t i c l e i n f o a b s t r a c t

Article history: The present work describes measurements of the effective thermal conductivity of NCFI 24-124 foam, a
Received 18 January 2012 spray-on foam insulation used formerly on the Space Shuttle external fuel tank. A novel apparatus to
Received in revised form 6 November 2012 measure the effective thermal conductivity of rigid foam at temperatures ranging from 20 K to 300 K
Accepted 25 November 2012
was developed and used to study three samples of NCFI 24-124 foam insulation. In preparation for mea-
Available online 23 January 2013
surement, the foam samples were either treated with a uniquely designed moisture absorption apparatus
or different residual gases to study their impact on the effective thermal conductivity of the foam. The
Keywords:
resulting data are compared to other measurements and mathematical models reported in the literature.
Polyisocyanurate foam
Insulation
Ó 2013 Elsevier Ltd. All rights reserved.
Thermal conductivity
SOFI

1. Introduction range. Prior to measurement, some of the samples were treated


with different residual gases or water vapor to investigate their
Liquid hydrogen and liquid oxygen are commonly used as rock- effect on the thermal conductivity. The data are compared to mea-
et fuel and as a result the aerospace industry has a demand for high surements reported in the literature and to mathematical models
quality cryogenic insulation that can be applied to launch vehicles. developed to predict the thermal conductivity of porous media.
One example of an application for porous insulating media is the
thermal protection system of the external tanks of the Space
2. Theory background
Shuttle, where spray-on foam insulation (SOFI) has been used for
decades [1]. This rigid foam insulation was selected for Space
Thermal conductivity is an important measure of insulation
Shuttle application because it has a low thermal conductivity
performance. However, because nearly all cryogenic insulation is
(k = 0.02 W/m K at room temperature) and high strength to density
inhomogeneous, the thermal conductivity of such materials is dif-
ratio (8–9 kPa/kg/m3) [2].
ficult to predict. Fourier’s conduction law provides the definition
Extensive research has been performed on thermal transport in
for the thermal conductivity, k(T), of a solid body:
foam insulations, although the vast majority has been associated
with near room temperature applications in, for example, the q ¼ kðTÞrT ð1Þ
building industry. Low temperature applications are unique due
When the material in question is isotropic and the temperature
to the potential for moisture absorption and other condensable
gases affecting thermal transport. Mathematical models have been difference is small, the thermal conductivity can be averaged over
the temperature range, and the one dimensional heat conduction
proposed to predict heat transport through foam insulation, but
these models generally have not been tested at low temperatures. of a constant cross-section sample becomes:
As a result, more experimental data on the foams of interest sub- 
kðTÞA DT
jected to the anticipated operating conditions are required. Q¼ ð2Þ
L
Here we report thermal conductivity measurements using a no-
vel apparatus that has been developed specifically to study rigid where L is the length of the sample in the direction of heat transfer
foam samples at temperatures ranging from 20 K to 300 K [3]. and A is the cross sectional area (perpendicular to the heat transfer
The effective thermal conductivity of three samples of NCFI 
direction). Note that in the Eq. (2), kðTÞ refers to the average thermal
24-124 foam insulation was measured over the full temperature conductivity of a solid. In the case of cryogenic insulation (e.g. MLI,
foam, aerogel beads), the conditions mentioned above (isotropic
⇑ Corresponding author. Address: Facility for Rare Isotope Beams, Michigan State and small DT) are rarely met.
University, USA. Tel.: +1 757 269 7058. In practice for anisotropic materials, it is normal to measure the
E-mail addresses: barrios@frib.msu.edu, barrios@jlab.org (M. Barrios). effective thermal conductivity, keff, of a sample. The effective

0011-2275/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.cryogenics.2012.11.004
M. Barrios, S.W. Van Sciver / Cryogenics 55–56 (2013) 12–19 13

Nomenclature

q heat flux CV constant volume specific heat


Q heat energy c average molecular speed
k thermal conductivity j Boltzmann constant
A cross sectional area a accommodation coefficient
T temperature i,j indices
l length of material in the heat transfer direction, cell Aij empirical function for use with gas mixtures
dimension n total number of gases in a mixture, refractive index
keff effective thermal conductivity y mole fraction of a gas in a mixture
ks effective solid conductivity e constant used in the Maxon–Saxena relation
kg effective gas conductivity P porosity
kr effective radiative conductivity km thermal conductivity of a solid matrix
Kn Knudsen number fs fraction of solid in the struts of a porous material
k mean free molecular path b extinction coefficient
L characteristic spacing of a medium bs extinction coefficient of solid polymer
p pressure qf foam density
M molecular mass qs solid polymer density
q density


thermal conductivity can replace kðTÞ in Eq. (2), but is assumed to the thermal conductivity of the solid matrix by subtracting the cal-
include additional heat transfer modes such as gas conduction, culated radiation contribution from effective thermal conductivity
radiation heat exchange, and in some cases convection and contact of the porous medium measured in vacuum. This latter method
resistance. may be reasonable for open cell foams, but for closed cell foams,
Mathematical models exist that estimate these components of performing measurements in vacuum does not ensure that the
keff under different conditions. Most models found in the literature residual gas has been entirely removed from the cells.
combine solid and gas thermal conductivity using geometric sim- Obtaining the gaseous contribution to the thermal conductivity,
plification of the internal structure of the foam [4–10]. Radiation kg, can also be difficult because it is dependent on the composition
is assumed to occur in parallel with the conduction modes and is and pressure of the contained gas. For most gases at pressures
added to the solid and gas conductivity to arrive at an estimate above about 1 Pa, the thermal conductivity is mainly a function
for the overall effective thermal conductivity. In the following sec- of temperature. However, at low pressures (p < 1 Pa) and for many
tion, we review various attempts to model keff for porous media. types of porous media, the mean free path of the gas molecules be-
The purpose is to provide basic understanding of the mechanisms comes larger than the pore size. In this case, the gas thermal con-
and to show where additional theoretical effort is necessary. ductivity is mainly function of the accommodation coefficient, a, a
quantity that determines how well the gas molecules transfer en-
2.1. Solid and gas conductivity ergy to the solid material. For this reason, it is most desirable to
know the gaseous pressure and temperature within the porous
Maxwell [7] was one of the first to examine conduction in het- media even though this may not be achievable for closed cell
erogeneous media. He developed a fairly simple expression for the foams.
electrical conductivity of a material consisting of spherical inclu- Springer [13] and Tien and Cunnington [14] considered four dif-
sions within a medium. This model can be adapted to describe ferent regimes of gaseous heat conduction based on the value of
the effective thermal conductivity as, the Knudsen number:
 
2kg þ ks þ Pðkg  ks Þ k
keff ¼ ks ð3Þ Kn ¼ ð4Þ
2kg þ ks  2Pðkg  ks Þ l

where ks is the thermal conductivity of the solid matrix material where k is the mean free path of the molecules and l is the pore size.
and kg is gas conductivity. The porosity (P) is defined as the ratio The four gas regimes are: free-molecule (Kn > 10), transition
of the pore volume to the total volume of the material. Use of Eq. (10 > Kn > 0.1), temperature-jump (slip) (0.1 > Kn > 0.01), and con-
(3) depends on knowledge of the thermal conductivity of the solid tinuum (Kn < 0.01). Springer estimated the gaseous heat conduction
matrix and contained gas and it ignores any contribution due to for the above regimes for simple geometric configurations (parallel
radiation heat transport. plates, coaxial cylinders, and concentric spheres) and small bound-
Determining the solid conductivity contribution to the effective ary temperature differences.
thermal conductivity requires knowledge of the thermal conduc- From kinetic theory the thermal conductivity of a gas in the
tivity of the solid matrix material, ks. If the thermal conductivity continuum approximation can be expressed by:
of the solid polymer that makes up the polyurethane (PU) or polyi- 1
socyanurate (PI) foam is known for the temperature range in ques- kg / cqC V k ð5Þ
3
tion it can be used in conjunction with the internal geometry to
determine the effective contribution from the solid matrix, km. where q is density, CV is the heat capacity at constant volume (J/
However, for some porous media, there is little data available for kg K), and c is the average molecular speed:
the thermal conductivity of the solid material. Tseng et al. [11] ap- rffiffiffiffiffiffiffiffiffiffiffi
8kB T
proached this problem by assuming the thermal conductivity of PU c ¼ ð6Þ
foam to be equal to that of nylon, a similar material. Still this ap-
pM
proach is limited by the incomplete knowledge of the internal where kB is the Boltzmann constant. However, for the residual gases
structure of the PU material. Alternatively, Wu et al. [12] estimated present in porous insulating media the small voids can cause the
14 M. Barrios, S.W. Van Sciver / Cryogenics 55–56 (2013) 12–19

gas to enter the free-molecule regime at low pressures. For such 16n2 rT 3
complex void geometries, Tien and Cunnington [14] suggested the kr ¼ ð11Þ
3b
use of an effective mean free path to develop the following empir-
ical relation for effective gas thermal conductivity in the transition where n is the refractive index of the porous media, r is the Stefan–
and temperature-jump regimes: Boltzmann constant, and b is the extinction coefficient, the inverse
0 of the mean penetration distance of radiation in the medium. Unfor-
kg ¼ akg ½l=ðl þ kÞ ð7Þ tunately, the extinction coefficient is another property that can vary
In the free-molecule regime, the apparent thermal conductivity with polymer type, morphology, and temperature. Glicksman [10]
can be estimated using Knudsen’s formula: analyzed the structures of cell walls and struts to find the following
relation for the extinction coefficient for PU and PI foams:
 rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
cþ1 R rffiffiffiffiffiffiffiffiffiffiffi
 ffi
kFM ¼ a  plDT ð8Þ fs q f
c1 8pMT  
qs ð1  fs Þqf
A further complication in the calculation of thermal conductiv- b ¼ 4:1 þ bs ð12Þ
l qs
ity occurs when there is a mixture of gases present. In many cases
when calculating the thermal conductivity of a mixture of gases, a where qf is the density of the foam, qs is the density of the solid
simple molar average of the thermal conductivity of the compo- polymer, fs is the fraction of solid in the struts of the foam, and bs
nents is insufficient [15]. In mixtures involving two gases with dif- is the extinction coefficient of the solid polymer. The first term on
ferent molecular weights, however, the thermal conductivity tends the right side of Eq. (12) is the contribution to the extinction coef-
to be lower than expected. The Wassiljewa equation can be used to ficient due to the struts. Because the struts are typically much thick-
account for this correction: er than the cell walls, they can be considered opaque. Therefore, the
X
n extinction coefficient of the solid material is not included in the
y ki
kg ¼ Pn i ð9Þ strut contribution.
i¼1 j¼1 yj Aij Although there has been a considerable effort applied toward
modeling thermal transport in porous media containing gases,
where kg is the thermal conductivity of the gas mixture, ki is the
the results are qualitative at best. In the present case of SOFI mate-
thermal conductivity of gaseous component i, n is the total number
rial at low temperature, the effective thermal conductivity is com-
of gases in the mixture, yi and yj are the mole fractions of compo-
plex because the internal structure of the material is not well
nents i and j, and Aij is a function to be specified (Aii = 1). For non-
characterized, the thermal conductivity of the solid matrix mate-
polar gas mixtures, Poling et al. [15] suggest the Mason–Saxena
rial is not well known and the material is a closed cell foam, with
relation:
the gas composition and pressure within the foam cells not well
h i1=2
e 1 þ ðki =kj Þ1=2 ðMi =Mj Þ1=4 known. One additional complexity comes about as a result of the
Aij ¼ ð10Þ low temperature application of SOFI. That is, the possibility that
½8ð1 þ Mi =Mj Þ1=2 in its application some of the contained gas may condense at low
temperature further complicating the effort to calculate the gas
where M is molecular weight of the gaseous components and e is a
contribution to the effective thermal conductivity. All these factors
constant near unity.
combine to make prediction of the effective thermal conductivity
Once values for ks and kg have been determined, it is then nec-
very challenging and thus demanding experimental measurements
essary to account for the tortuous path that heat conduction fol-
to properly characterize the material property.
lows within the structure of the porous media. This is especially
important in evacuated media, when gas conduction is minimized.
Several heat transfer models [4–6] developed with heat conduction 3. Thermal conductivity measurement apparatus
through loose fill insulations can be simplified for use with rigid
foam calculations. These models involve the additional complexi- The experimental apparatus developed for measurement of the
ties of surface contact, packing arrangement, and porosities on dif- effective thermal conductivity of flat plate samples of SOFI or other
ferent scales (i.e. hollow glass spheres, aerogel beads) which can solid materials at low temperatures has been described previously
generally be neglected in the case of rigid foam. Other models have [3]. Here we provide a summary of the important characteristics of
been developed specifically for rigid foam [7–10]. this device so that the reader can best understand the experimen-
Many approaches found in the literature analyze a unit cell tal methods and results.
within the microstructure of the porous media. The unit cell is de- The ASTM standard C177, ‘‘Standard Test Method for Steady-
fined by Dul’nev [4] as the smallest volume whose effective heat State Heat Flux Measurements and Thermal Transmission Proper-
conductivity coincides with the effective heat conductivity of the ties by Means of the Guarded-Hot-Plate Apparatus,’’ [17] was used
disperse system. The unit cell can be represented in many forms: as a template and modified to accommodate the additional
cubic inclusion, spherical inclusion, cubic skeleton with intercon- requirements for operation in a cryogenic environment. The appa-
necting pores, etc. Different methods are used to find the thermal ratus design uses a single-sided guarded-hot-plate with only one
conductivity of the unit cell chosen to represent the internal struc- cold plate and one specimen. This approach greatly simplified
ture of the system. Geometric parameters are then used to relate the design of the cooling plate, which was connected directly to
the dimensions of the unit cell to the porosity of the media. The a cryocooler as a heat sink. Also, a double-sided design would re-
effective thermal conductivity of the media is then found as a func- quire a more complicated thermal link and thermal stabilization
tion of the solid and gas conductivities and the porosity. between the two cold sides. Fig. 1 is a schematic of the experimen-
tal chamber.
2.2. Radiative conductivity The apparatus was calibrated using a disk of nylon with a
known thermal conductivity as a control sample. Through this pro-
Radiation heat transfer can significantly contribute to the effec- cess, it was found that contact resistance between the faces of the
tive thermal conductivity of a porous medium. For optically thick sample and the hot and cold plates was a limiting factor, even
media, Siegel and Howell [16] suggested an effective radiative con- when copper grease was applied to the faces of the sample. The
ductivity given by: thickness (25.4 mm) of the samples and the diameters of the
M. Barrios, S.W. Van Sciver / Cryogenics 55–56 (2013) 12–19 15

To Cryocooler

Stainless Steel Plate Copper Stem

Hot Cold Plate


Plate Insulation Sample
Phenolic
Guard

Compression Springs Aluminum Guard

Vacuum Chamber

Fig. 1. A schematic of the experimental chamber for measurement of the effective thermal conductivity of solid foam samples.

temperature sensors (3.175 mm) created a large enough uncer- head of the cryocooler to control the overall temperature of the
tainty in the gauge length to eliminate the possibility of embed- apparatus. Stainless steel tubes are welded to the top of the cold
ding the sensors within the sample. It was thereafter decided plate to provide mechanical supports. The apparatus is placed in-
that, in order to minimize the contact resistance, the foam sample side an evacuated cryostat with a liquid nitrogen shield to isolate
should be bonded to copper plates using Stycast 2850 epoxy. The it from ambient. Fig. 2 is a schematic of the entire apparatus.
Stycast provides excellent thermal contact with the faces of the The effective thermal conductivity of the SOFI samples was
sample. Temperature sensors were then mounted to the copper measured by recording the temperature on either side of the sam-
plates. The added thermal resistance of the Stycast layers and cop- ple for a specific heater power after steady state was achieved.
per plates was determined to comprise less than 1% of the overall Prior to installing a sample in the apparatus, the diameter and
sample resistance. thickness were measured at various locations. These measure-
The experimental cell is attached to a cryocooler (Cryomech ments were then used to calculate the surface area and the average
model PT-810) at the top of the cold plate by means of a copper thickness of the sample. The heater power was measured by
stem that penetrates the stainless steel plate and a braided copper recording the voltage across the heater and the voltage across a
thermal link. The cryocooler is capable of providing 14 W of cool- known resistor in series at room temperature to yield the current
ing power to the apparatus at 20 K. A heater is mounted to the through the circuit. The temperatures were measured using Cernox

Fig. 2. A schematic of the apparatus for measurement of the effective thermal conductivity of SOFI.
16 M. Barrios, S.W. Van Sciver / Cryogenics 55–56 (2013) 12–19

sensors on the copper plates attached to the specimens and on the


aluminum guard.
Before calculating the effective thermal conductivity, a correc-
tion was made to account for any heat leak between the hot plate
and aluminum guard. The typical accuracy of the temperature sen-
sors was ±40 mK. This translates to an uncertainty of less than 1%
in temperature measurement. The heater voltage resolution on the
Lakeshore 340 temperature controller is 1.25 mV. The estimated
minimum required operating voltage is 553 mV. This corresponds
to a maximum heating power uncertainty of 0.2%. Width and area
measurement uncertainty were limited to below one percent sim-
ply by performing several measurements with a micrometer. Ther-
mal contraction measurements from the literature vary, but do not
exceed 2% between room temperature and 80 K. The sum of the er-
rors listed above is reflected in the overall 6% error bars seen in the
results. For most measurements a DT of 10 K across the sample
was chosen to make it possible to determine the effective thermal
conductivity with an instrument error below 5%.
Fig. 3. The thermal conductivity versus average temperature for NCFI 24-124
sample #1 in the ‘‘as received’’ and evacuated cases compared with literature data
4. Measurements and discussion [18]. The increase at low temperature in the ‘‘as received’’ case is probably the
result of residual helium gas. The error bars correspond to a 6% uncertainty
The thermal conductivity of three samples of NCFI 24-124 foam determined by the temperature sensors, heater, sample width and area measure-
ment, and thermal contraction.
was measured. This foam has a porosity of 97% and an average pore
diameter of 0.3 mm. Table 1 lists the three samples and shows the
conditions under which the thermal conductivity was measured. controlled area for 18 months. Our sample was aged in the same
The first NCFI sample (#1) was tested ‘‘as received’’ by installing area for 18 months, and stored in a sealed plastic bag for 4–5 years
it within the experimental chamber under room temperature and before testing. The discrepancy in the data could be due to aging,
atmospheric pressure of air (296 K and 102.6 kPa) and sealing the diffusion of the low thermal conductivity blowing agent out of
chamber. The thermal conductivity was then measured at various the foam over time, or the presence of the higher thermal conduc-
temperatures between 296 K and 30 K. The gas pressure in the tivity helium gas.
experimental chamber varied from 102.6 kPa at 296 K to For the vacuum case, sample #1 was placed in the experimental
0.267 kPa at 30 K. Subsequently, in order to examine the effect of chamber and evacuated, using a turbo-molecular vacuum pump,
different residual gases on the thermal conductivity of the foam, for one week at room temperature. At this point the pressure inside
sample #1 was tested under vacuum and with helium gas added the chamber was 138.3 Pa, decreasing at less than 0.4 Pa per hour.
to the experimental chamber. The cryocooler was then turned on and the sample cooled to 30 K.
Fig. 3 shows the thermal conductivity of sample #1 in air and As expected, the measurements show a clear decrease in effective
under vacuum. The thermal conductivity shows a clear decrease thermal conductivity compared to the as received case with helium
with temperature down to 80 K followed by a sharp increase. As contamination. The data for the evacuated case follow a similar
is discussed below, this increase in keff probably results from the trend to that of the as received case at high temperature, but do
condensation of the lower thermal conductivity heavy gases onto not reproduce the anomaly below 80 K.
the cell walls thus increasing the molar ratio of higher thermal In order to further examine the effect of residual gas on the
conductivity trace gases such as helium. Tseng et al. [11] observed thermal conductivity of the foam, sample #1 was placed in the
a similar phenomenon for newly sprayed polyurethane foam be- experimental chamber, evacuated overnight and then purged with
tween 230 K and 280 K. In this temperature range, the R141b gas 101 kPa of helium gas at ambient temperature three times. After
present in the cells condenses and the air remaining in the cells the third filling, the cryocooler was then turned on and the thermal
dominates the thermal transport. In the present case, helium was conductivity measurement repeated. Due to the 97% high porosity
inadvertently introduced into the foam during a leak checking pro- of the foam, the residual gas can dominate the effective thermal
cess. As a result during the measurement, the air present in the conductivity. This is shown in Fig. 4, where the effective thermal
cells condensed below 80 K and the remaining helium gas domi- conductivity of the foam with helium as a residual gas is compared
nated the heat transport dramatically increased the effective ther- to that of the evacuated and as received cases. The effective ther-
mal conductivity of the SOFI sample. mal conductivity of the helium case is roughly three times larger
The data from Fesmire et al. [18], shown as the solid circles in than that of the as received case. In this regime, the thermal con-
Fig. 3, are the effective conductivity of the same material as our ductivity of helium gas is at least six times larger than that of air
sample, recorded at average temperatures of 297 K and 185 K, and the thermal conductivity of the solid material is greater than
respectively. Their data appears slightly below that found in the that of helium gas, so it is surprising that the effective thermal
present study. Fesmire et al. measured a sample aged in a climate conductivity measured is lower than that of pure helium gas. This

Table 1
NCFI 24-124 samples and the conditions under which their thermal conductivity was measured. Sample #1 (‘‘as received’’, x) was subjected to an unknown amount of helium gas
during the leak check process, which is believed to have affected the thermal conductivity.

Sample number As received Water vapor conditioned Helium purged Under vacuum Helium/air mixtures
#1 x x x
#2 x x x
#3 x
M. Barrios, S.W. Van Sciver / Cryogenics 55–56 (2013) 12–19 17

Fig. 5. The thermal conductivity vs. temperature for sample #2 with various
Fig. 4. The thermal conductivity data of NCFI 24-124 sample #1 for the helium residual gas mixtures. Lines through the data are guides to the eye.
purged case, the ‘‘as received’’ case, and the evacuated case. The thermal
conductivity of the conditioned sample #2 ( ) is shown for comparison.

of the molecules is much shorter than the characteristic dimension


discrepancy is almost certainly due to insufficient purging of the of the cells within the foam.
sample. Because of the closed cell nature of the foam it is very dif- Fig. 5 shows that the thermal conductivity of the 85.3 kPa of air
ficult to replace all of the R141b gas and air originally present in case for sample #2 follows the thermal conductivity of the condi-
the cells with helium gas. In any case, the trend in the thermal con- tioned case closely until around 80 K where it begins to approach
ductivity of the helium purged case approaches that of the sample the thermal conductivity of the evacuated case. This is expected
#1 ‘‘as received’’ case below 50 K. This is further evidence that in as uncontaminated air will condense and freeze below 80 K and
the as received case, the sample was contaminated with helium leave a near vacuum environment within the cells.
gas. The presence of 0.3% helium gas should cause the thermal con-
Also seen in Fig. 4 are the data from the conditioned NCFI ductivity of the sample to be larger than that of the air and evacu-
24-124 sample #2. Sample #2 was tested in the same fashion as ated cases at temperatures below 80 K. It can be seen in Fig. 5 that,
sample #1; but it was first subjected to a launch pad conditioning in fact, the thermal conductivity begins to increase as the temper-
process. The conditioning process subjected the foam to a typical ature is lowered below 100 K. This marked increase shows that
environment expected on the launch pad in Cape Canaveral even a small amount of helium gas produces an increase in the
(T = 34 ± 2 °C, Relative Humidity >75%) for 8 h. This conditioning effective thermal conductivity of the foam at lower temperatures.
process causes moisture absorption in the foam, potentially affect- As expected, the thermal conductivity of the sample was much
ing its thermal properties [19]. Above 70 K, the thermal conductiv- higher due to the higher mole fraction of helium (see Fig. 5). How-
ity of the conditioned sample is nearly the same as the ‘‘as ever, as the temperature of the sample increased from 40 K to 60 K
received’’ sample. This shows that the conditioning process and an unexpected drop in effective thermal conductivity was ob-
moisture absorption have little effect on the thermal conductivity served. Near the observed pressure at 40 K, the sublimation line
of the foam. nitrogen is crossed. This could cause some of the nitrogen to leave
To quantify the effect of helium intrusion on the thermal con- the surfaces of the foam. As the nitrogen pressure increases, it
ductivity of the foam at low temperature, sample #2 was first evac- mixes with the helium gas and impedes the overall heat transport,
uated for several weeks and the thermal conductivity then thus lowering the effective thermal conductivity. However, a corre-
measured between 30 K and 290 K. This provided a baseline for sponding increase in pressure is not recorded in the measurements
the thermal conductivity of the foam with minimal contribution possibly because the pressure increase is small and the time neces-
from the residual gas. The chamber was then filled with air to a sary for the gas to diffuse out of the sample had not elapsed.
pressure of 85.3 kPa. A period of 24 h was then given for the air The sample #2 effective thermal conductivity measurements
to diffuse into the foam during which time additional air was were then compared to the Maxwell model, Eq. (3). This model re-
added to the chamber to maintain the pressure constant. The ther- quires knowledge of the porosity and thermal conductivity of the
mal conductivity of the sample was then measured from 20 K to solid material and the gas. Values for the porosity and solid
120 K. thermal conductivity at room temperature were taken from the
The sample was then warmed to near room temperature and a literature [20,21]. Below room temperature, the solid thermal con-
small amount (267 Pa partial pressure) of helium gas was added to ductivity was extrapolated by reproducing the relationship
the 85.3 kPa of air. This brought the mole fraction of helium gas to between temperature and thermal conductivity seen in similar
0.3% at room temperature. At low temperature the helium gas materials, such as PVC and nylon. The cell dimensions were deter-
would then be in the transition gas conduction regime. Lastly, sam- mined from average values measured using SEM images [19]. The
ple #2 was warmed to room temperature and more helium was residual gas thermal conductivity, required in all of the models, is a
added up to a partial pressure of 16 kPa, which brought the total difficult parameter to estimate. In order to do so accurately, the
gas pressure to 101 kPa at room temperature and the mole fraction mixture of gases and their partial pressures must be known. Also,
of helium to 16%. This mixture was chosen to maintain the contin- at low temperatures the condensation of gases must be taken into
uum gas regime over the entire temperature range 20–300 K. As account (e.g. in a mixture of nitrogen and helium, the nitrogen will
discussed previously, the continuum regime occurs when the condense at higher temperatures than helium, causing the molar
Knudsen number is less than 0.01, meaning the mean free path ratio of helium in the mixture to increase). At low pressures, as
18 M. Barrios, S.W. Van Sciver / Cryogenics 55–56 (2013) 12–19

Fig. 6. The effective thermal conductivity data for the evacuated NCFI 24-124
sample #2 compared to predictions based on the Maxwell model, Eq. (3). Fig. 8. The thermal conductivity data for the NCFI 24-124 sample #2 with an initial
mixture of 85.3 kPa air and 16.0 kPa helium (84% Air/16% He) at 300 K plotted
against predictions from the Maxwell model, Eq. (3).

Fig. 7. The thermal conductivity data for the NCFI 24-124 sample #2 with air as a
residual gas at 85.3 kPa at 300 K plotted against predictions from the Maxwell
model, Eq. (3). Fig. 9. The effective thermal conductivity for two NCFI 24-124 samples. Sample #2
is launch pad conditioned and sample #3 as received.

continuum gas heat transfer transitions into the Knudsen regime,


the gas thermal conductivity becomes even more difficult to esti-
Comparisons between the effective thermal conductivity results
mate. The mathematical models do not take these factors into ac-
and the Maxwell’s model make it clear that in order to make accu-
count, and it is left to the analyst to determine the appropriate
rate predictions of keff in SOFI it is necessary to know the types of
value for gas thermal conductivity at each point.
residual gases present in the foam. Furthermore, it is important
Fig. 6 is a comparison between Maxwell’s model and the data
to understand how the pressures of the gases change with temper-
from the evacuated case for sample #2. The model tends to under
ature. The large surface areas present in the foam can cause gases
predict the thermal conductivity at temperatures above 40 K. The
to condense at temperatures higher than would normally be
sharp increase in thermal conductivity from 30 K to 50 K is possi-
expected.
bly due to sublimation of nitrogen inside the cells of the foam.
The effective thermal conductivity of a third NCFI 24-124 sam-
The nitrogen increases the thermal conductivity of the foam, but
ple #3 was measured ‘‘as received’’ to check reproducibility and
since the foam is closed cell the increased pressure inside the foam
obtain data without helium gas contamination. These results are
is not recorded by the pressure gauge, and is therefore not reflected
compared to the launch pad conditioned sample #2 in Fig. 9. As
in the model.
can be seen in the figure, the conditioning process has little effect
Fig. 7 is a comparison between the Maxwell model and the data
on the effective thermal conductivity of the foam resulting in at
from the 100% air sample. Again, the model fails to predict the in-
most a 10% increase.
crease in thermal conductivity between 35 K and 55 K. This is a fur-
ther indication that the gauge pressure does not accurately depict
the pressure inside the cells at low temperatures. 5. Conclusions
Fig. 8 compares the measurements of the 84% air/16% He sam-
ple to the model. In this case, the model predicts the increase in The goal of this work was to measure the effective thermal con-
thermal conductivity from 20 K to 40 K, but greatly over predicts ductivity of flat plate samples of spray-on foam insulation (SOFI)
the conductivity from 50 K to 110 K. Again, this is due to fluctuat- with different preparation and at temperatures ranging from
ing gas pressures inside the closed cells, causing the thermal con- 20 K to 300 K. A single sided guarded-hot-plate apparatus was
ductivity to decrease as nitrogen sublimates. developed for this purpose.
M. Barrios, S.W. Van Sciver / Cryogenics 55–56 (2013) 12–19 19

For sample #1 as received, the data show a clear decrease in References


thermal conductivity with decreasing gas pressure in the experi-
mental chamber with the data for the evacuated case following a [1] Harvey J, Butler J, Chartoff R. Development of polyisocyanurate pour foam
formulation for Space Shuttle external tank thermal protection system. NASA
similar trend to that of the as received sample. In any case, the technical report. Marshall Space Flight Center; August 1988.
trend in an anomaly in the effective thermal conductivity below [2] Sparks LL. Low-temperature properties of expanded polyurethane and
50 K appears to have been caused by helium gas contamination. polystyrene. In: McElroy DL, Tye RP, editors. Thermal insulation
performance, ASTM STP 718, American Society for Testing and Materials;
The effective thermal conductivity of the sample #2 was then 1980. p. 431–52.
studied to explore the effect of launch pad conditioning and helium [3] Barrios M, Van Sciver SW. An apparatus to measure thermal conductivity of
gas contamination. The conductivity was first measured to deter- spray-on foam insulation. In: AIP conference proceedings, vol. 1218; 2010. p.
938–45.Barrios M, Vanderlaan and Van Sciver SW. Thermal Conductivity of
mine the effect of conditioning. Following this measurement, the
Spray-On Foam Insulations for Aerospace Applications. AIP Conference
sample was first subjected to 85.3 kPa of air at 300 K and its the Proceedings, Vol. 1434, pp. 1319-1326, 2012.
thermal conductivity closely followed the previous measurements [4] Dul’nev GN. Heat transfer through solid dispersed systems. Eng Phys J
down to 80 K where it approached the thermal conductivity of the 1965;9:275.
[5] Odelevskii VI. Calculation of the generalized conductivity of heterogeneous
evacuated case. This is expected because as the temperature de- systems. Zh Tekh Fiz 1951;21(6):667.
creases, the air will adhere to the surfaces of the foam and leave [6] Luikov AV, Shashkov AG, Vasiliev LL, Fraiman YE. Thermal conductivity of
a vacuum. When a partial pressure of 0.267 kPa of helium was porous systems. Int J Heat Mass Transfer 1968;11:117–40.
[7] Maxwell JC. A treatise on electricity and magnetism. 3rd ed. vol. 1. Oxford:
added to the 85.3 kPa of air at room temperature, the effective Clarendon Press; 1892. p. 440.
thermal conductivity increased above the base case below 100 K. [8] Russell HW. Principles of heat flow in porous insulators. J Am Ceram Soc
This marked increase shows that even this small amount of helium 1935;18:1–5.
[9] Schuetz MA, Glicksman LR. A basic study of heat transfer through foam
gas affects the thermal conductivity of the foam at lower temper- insulation. J Cell Plast 1984;20(2):114–21.
atures. As more helium is added to the residual gas mixture an in- [10] Glicksman. Heat transfer in foams. In: Hilyard NC, Cunningham A, editors. Low
crease in thermal conductivity can be seen over the entire density thermal plastics. London: Chapman and Hall; 1994.
[11] Tseng C, Yamaguchi M, Ohmori T. Thermal conductivity of polyurethane foams
temperature range. Measurements below 120 K show that the from room temperature to 20 K. Cryogenics 1997;37(6).
effective thermal conductivity does not correlate with Maxwell’s [12] Wu J, Sung W, Chu H. Thermal conductivity of polyurethane foams. Int J Heat
model. This is probably due to the inability to know the partial Mass Transfer 1999;42:2211–7.
[13] Springer GS. Heat transfer in rarefied gases. In: Irvine TF, Hartnett JP (editors).
pressures of the gas mixtures within the cells at low temperature.
Advances in heat transfer. vol. 7; 1971. p. 163–218.
These effective thermal conductivity results stress the impor- [14] Tien CL, Cunnington GR. Cryogenic insulation heat transfer. In: Irvine Jr TF,
tance of residual gases present in foams like SOFI. Although launch Hartnett JP, editors. Advances in heat transfer, vol. 9. New York: Academic
pad conditions may slightly increase the thermal conductivity of Press; 1973. p. 349–417.
[15] Poling B, Prausnitz J, O’Connell J. The properties of gases and liquids. New
the foam, exposure to high thermal conductivity gases such as he- York: McGraw Hill; 2001.
lium or hydrogen can have a much greater effect. In the case of [16] Siegel R, Howell JR. Thermal radiation heat transfer. 2nd ed. McGraw-Hill book
hydrogen storage, exposing the foam to venting or leaking hydro- Company; 1981.
[17] ASTM C177. Standard test method for steady-state heat flux measurements
gen gas can cause a significant increase in the effective thermal and thermal transmission properties by means of the guarded-hot-plate
conductivity of the foam. apparatus. Annual book of ASTM standards; 04.06, (2000a).
[18] Fesmire JE, Coffman BE, Meneghelli BJ, Heckle KW. Spray-on foam insulations
for launch vehicle cryogenic tanks. Cryogenics; January 26, 2012.
[19] Barrios M. Material characterization of rigid foam for aerospace vehicles. PhD.
Acknowledgements Dissertation, Florida State University, Department of Mechanical Engineering;
2011.
This work was supported by NASA Kennedy Space Center and [20] Sullivan RM, Ghosn LJ, Lerch BA. Application of an elongated Kelvin model to
Space Shuttle Foams. J Spacecraft Rockets 2009;46(2):411–8.
the Florida Center for Advanced Aero-Propulsion (FCAAP). The Na- [21] Stewart J. The influence of morphology on polyurethane foam heat transfer.
tional High Magnetic Field Laboratory is supported by the NSF and Masters thesis, Massachusetts Institute of Technology, Department of
the State of Florida. Mechanical Engineering; 1994.

S-ar putea să vă placă și