You are on page 1of 6

Ind. Eng. Chem. Res.

1993, 32, 3117-3122 3117

Supercritical Fuel Deposition Mechanisms


Tim Edwards*
USAF Wright Laboratory, WL/POSF Bldg 490, 1790 Loop Rd N, Wright-Patterson AFB, Ohio 45433-7103

Steven Zabarnick
300 College Park, KL463, University of Dayton Research Institute, Dayton, Ohio 45469-0140

Jet fuel is the primary coolant used in high-speed aircraft. A decrease in surface deposition (fouling)
is often seen as jet fuels and pure hydrocarbons are heated above approximately 370 °C, as measured
by the wetted wall or film temperature. This temperature is near the critical temperature of most
jet fuels. Two explanations have been offered for this decrease in deposition under supercritical
conditions. One explanation is that the decrease reflects the temperature where hydroperoxide
precursors to solids formation are depleted by thermal decomposition, interrupting the radical
chain reactions forming solids. Another explanation is that the solvent properties of the fuel become
enhanced under supercritical conditions, with the surface deposition reduced essentially by keeping
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

the solids in solution. In single-tube heat exchanger tests with pure hydrocarbons and jet fuels of
widely varying critical temperature, it was found that the deposition decrease is insensitive to fuel
critical temperature but very sensitive to residence time/heating rate, indicating that the deposition
decrease is a fuel chemistry effect rather than an effect of the supercritical nature of the fuel.
Downloaded via ONERA on March 7, 2023 at 09:07:05 (UTC).

Introduction Table I. Critical Properties of Current Jet Fuels (Martel,


1978) and n-Paraffins for Reference*
Development of thermally stable jet fuels has been TC,8F O Pc, psia Pc, atm
recognized as a critical research area because an aircraft’s
fuel is the only practical means of absorbing dramatically JP-4 (Cg.sHie.s) 550-690 288-365 500-400 34.0-27.2
JP-5 (C11.9H22.2) 720-780 382-415 300-280 20.4-19.0
increasing on-board heat loads. The thermal stability of JP-7 (C12.1H24.4) 760 405 270 18.4
a fuel is a measure of its tendency to cause fuel system JP-8 (C10.9H20.9) 700-760 370-405 350-275 23.8-18.7
problems, such as surface deposits or filter plugging, upon n-CsHig 564 296 365 24.8
heating during aircraft operation (Hazlett, 1991). Indeed, 726 386 263 17.9
the limited thermal stability of current fuels must be n-CieH34 836 447 200 13.6
overcome in order to fully realize the system capabilities “
Critical temperatures for fuels were calculated from the specific
of advanced aircraft. The current practice for Air Force gravity and boiling curve of the fuels by the method outlined in
engine designers is to limit the fuel to 163 °C (325 °F) bulk ASTM D2889. Fuel critical properties are listed as a range because
temperature and 205 0C (400 0F) wetted wall temperature. of varying boiling curves for different samples. Commercial Jet A
This “rule-of-thumb” is based on engine fouling experience. (C11.6H22.0) and Jet A-l fuels are similar to JP-5 and JP-8. Fuel

In the long term, the heat management requirements of composition data from (Martel, 1988).
future Air Force aircraft are expected to be sufficiently
demanding that the aircraft fuel system will need to operate exposed to heat and fuel flow) is on the order of hours.
at much higher temperatures. The goal for this long term Most researchers have seen a peak in surface deposition
near 370 °C (700 °F) when the deposition is plotted as a
fuel is to increase the thermal stability temperature limit
to480°C (900°F) (Edwards, 1993). This high temperature function of wetted wall or film temperature. This is
is above the fuel thermodynamic critical temperature, illustrated in Figure 1 for two sets of data from different
which typically ranges from 370 to 415 °C (700-780 °F) test devices (Taylor, 1974; TeVelde and Glickstein, 1983).
for current jet fuels (Martel, 1978). Typical fuel system The peak disappears if the dissolved oxygen in the fuel,
pressures are approximately 34-68 atm (500-1000 psia), typically about 65-75 ppm (w/w) in air-saturated fuel, is
well above the fuel critical pressure. Calculated critical removed (Bradley et al., 1974; Taylor, 1974). One goal of
data for typical fuels are shown in Table I. The physical our work is to investigate the source of this deposition and
and chemical properties of supercritical fuels are expected determine ways to reduce, avoid, or eliminate it without
to be dramatically different from gases and liquids. There resorting to removing the dissolved oxygen from the fuel.
is little information available on the properties of fuels The decrease in deposition above approximately 370
when operated at supercritical conditions (Edwards, 1993). °C has been given both physical and chemical explanations.
One area of concern for fuels operating at high tem- According to Taylor (1974), “the sharp drop in autoxidative
peratures is their thermal decomposition and the formation deposit formation rates... would appear to reflect an effect
of insoluble products that deposit onto metal surfaces of the transition between the liquid phase and the
(fouling). This area has been studied in heated metal tubes supercritical vapor phase.” Hazlett (1977) prefers a
(simulating aircraft nozzles and heat exchangers) by chemical explanation: “The deposit rating decreases above
various investigators (Hazlett, 1991; Hazlett et al., 1977; 385 °C. Since hydroperoxide is depleted at this temper-
Bradley et al., 1974; TeVelde and Glickstein, 1983; ature, the tie-in between hydroperoxide decomposition
Marteney and Spadaccini, 1986; Taylor, 1974; Faith et al., and deposit formation is reinforced.” Similarly, Marteney
1971; Chin and Lefebvre, 1992). These high-temperature and Spadaccini (1986) state “the sudden decrease in
tests cover a wide range of flow rates and residence times. deposit formation at temperatures above 645 K [372 °C]
The residence time of a given fuel molecule inside the may be indicative of a depletion of oxygen in the fuel,
heated tube is typically on the order of seconds, while the related to the known decomposition of hydroperoxides at
test time (or the amount of time the metal surface is elevated temperatures.”
0888-5885/93/2632-3117$04.00/0 © 1993 American Chemical Society
3118 Ind. Eng. Chem. Res., Vol. 32, No. 12, 1993

Figure 1. Surface deposition data for JP-5 in UTRC single-tube


heat exchanger test (TeVelde and Glickstein, 1983) and in Esso
Advanced Kinetic Unit (AKU) (Taylor, 1974). UTRC conditions: 28
atm, 570 mL/min (60 lb/h, inlet Re = 3000), 2.4-m-long, 2.16-mm-i.d.
316 SS tube, 480 °C (900 °F) fuel outlet temperature, 19 h test
duration. AKU conditions: 69 atm, 10 mL/min (inlet Re = 64), 122
-cm-long, 2.1-mm-i.d. 304 SS tube, 4 h test duration.

The research described in this paper was motivated by assessment of particulates formed in the heated section.
a desire to discover if the decrease in surface deposition A three-way valve downstream of the back pressure valve
above about 370 °C, which creates the deposition peak, is allows the product flow to be diverted to a sampling system
indeed due to supercritical fuel properties. Although other where the liquid products are analyzed (off-line) at ambient
conditions. The stainless steel test section is cut into 5-cm
investigators have studied a wide variety of conditions,
most ofthefuelsand hydrocarbons tested had very similar (2-in.) sections and analyzed by carbon burnoff (Leco
critical temperatures. One of the approaches of this work surface carbon analyzer) to measure carbon deposition.
was to investigate several similar hydrocarbons (e.g., the
The tube wall temperature distribution is measured by
normal paraffins) with widely varying critical tempera- K-type thermocouples spot-welded to the outside of the
tures, but presumably similar deposition chemistry, to tube. The fuel temperature at the furnace outlet is
determine the effect of fuel critical temperature on the measured by a thermocouple inserted into the flow.
Pressure was measured between the filter and the back
deposition behavior. A second approach was to test the
effect of fuel flow rate on the deposition-temperature pressure valve and also at the pump. The fuel temperature
behavior above 370 °C. The flow rate affects the wall/fuel distribution is estimated with a heat transfer program
temperature d ifference, the residence time, and the heating (Edwards, 1992; Edwards and Anderson, 1993) from the
rate. Pure hydrocarbon fuels are fairly expensive, so JP-7 measured wall temperature distribution, fuel flow rate,
was used in most of the tests as the baseline hydrocarbon
and the initial temperature of the fuel entering the furnace
fuel. JP-7 (MIL-T-38219) is the Air Force’s highest (20 °C). The fuel properties (density, thermal conduc-
thermal stability fuel and consists mainly of paraffins and tivity, viscosity, heat capacity) vary significantly through-
out the tube in a given test where the fuel temperature
cycloparaffins. Finally, a representative Jet A-l fuel was
studied for comparison to the JP-7 and pure hydrocarbon might vary from 20 to 620 °C (70-1150 °F) (Edwards,
results. These results will be used to develop computa- 1993), so the calculation uses temperature-dependent fuel
tional fluid dynamic and chemistry (CFDC) models of properties. It is found that test times on the order of
fuel system components (Katta and Roquemore, 1993). 10-20 h are required for JP-7 and pure hydrocarbons to
give reproducible deposition results above the measure-
ment background. For more typical (lower thermal
Experimental Section
stability) fuels, deposition amounts are much larger and
The apparatus used in this study is shown in Figure 2. test times can be shorter.
The fuel is air-sparged to assure air saturation of the fuel.
The flow system is initially purged with nitrogen prior to Results and Discussion
fuel introduction to remove any air present. The fuel is
pumped through the system with a high pressure liquid A typical result of a JP-7 experiment is shown in Figure
chromatography pump. An initial 0.45-/im filter is used 3. As described in the Introduction, the data is reduced
to remove any particulates present from fuel handling and by plotting the deposition (scaled by the tube surface area
to maintain a consistent filtration level between fuels. The and the total fuel flow) as a function of wall temperature,
test section consists of an 89-cm-long, 3.2 mm (0.125-in.) excluding points of decreasing wall temperature after the
o.d., 1.4-mm (0.055-in.) i.d. 304 stainless steel tube passing tube exits from the furnace. Thus, the deposition is
through a Lindbergh laboratory furnace. The tube makes reported as ng of deposit carbon/(g of fuel fed-cm2), or
a 180° bend inside the furnace, thus passing through the ppm/cm2. The deposition-wall temperature results for
30-cm (12-in.) actively heated portion of the furnace twice. several 12 mL/min flow rate experiments are shown in
After exiting from the furnace, the products are cooled to Figure 4. Various tests are shown spanning different
room temperature and filtered before passing through a ranges of wall temperature. The decrease in deposition
back pressure valve which regulates the system pressure. near 440 °C (825 °F) is very similar to the behavior shown
A 5-Mm filter is used to prevent particulates from fouling in Figure 1, although the deposition decrease begins at a
the back pressure valve; the filter also gives a visual temperature significantly higher than the fuel critical
Ind. Eng. Chem. Res., Vol. 32, No. 12,1993 3119
©— WalJT.C
S— Fuel T, C (calc) 0 C deposition, pg
-X— Fuel T, C (meas)

Figure 5. Results from 33.5 mL/min tests. Air-sparged JP-7 fuel,


69 atm.
Figure 3. Experimental results from 12.0 mL/min test. Air-sparged
JP-7 fuel, 69 atm, 21 h.

Figure 6. Results from 3.1 mL/min tests. Air-sparged JP-7 fuel, 69


atm.
Figure 4. Combined results from 12 mL/min tests covering various
wall temperature ranges. Air-sparged JP-7 fuel, 69 atm. Residence
times and Reynolds number of bulk fuel at peak deposit location
rates, the deposition-temperature behavior is more likely
a result of a chemical, rather than a physical (supercritical)
calculated with temperature-dependent fuel properties. Lines connect
points in order of axial distance. process.
It is notable in Figures 4-6 that the background
deposition levels are highest for the shortest test times.
temperature of 400 °C (760 °F). The dependence of this For short test times (< 10 h), it is found that the deposition
deposition decrease on residence time/flow rate was results at temperatures above and below the autooxidative
investigated by running similar tests at 33.5 and 3.1 mL/ peak were sufficiently close to background detection levels
min flow rates. The results of these tests are shown in that the data are fairly scattered. This can be seen in
Figures 5 and 6, respectively. It is immediately apparent Figure 4 (test 5/92-2) and Figure 5 (test 4/92-3). Apparent
that the autooxidation deposition decrease has shifted background levels varied somewhat for the various tests,
significantly with the varying flow rates and residence but typical levels were on the order of 7 fig/cm2. Recent
times. As fuel flow rates are increased, the difference tests with lower stability fuels have shown an induction
between the fuel and wall temperature increases, a simple time, or delay before the initiation of deposition, of about
3 h under conditions similar to Figure 4 (Edwards and
result of the increasing heat-transfer requirements to heat
Liberio, 1994).
up the increasing amount of fuel. Thus, the fuel tem- The results for the varying flow rates are summarized
perature at the peak has also changed. If the deposition in Table II. In addition to the change in deposition peak
decrease is a simple consequence of the enhanced solid
location, the deposition (expressed as ppm) decreases with
solubility of the fuel under supercritical conditions, the increasing flow rate. The only published results for varying
peaks in Figures 4-6 should occur at the same wall flow rates at high temperatures were those of UTRC
temperature (since the deposit would be dissolved by the (TeVelde and Glickstein, 1983; Marteney and Spadaccini,
fuel near the wall which should be at a temperature near 1986). UTRC examined a wider range of flow rates in
that of the wall). The bulk fuel temperature at the point their heated tube tests. For inlet Reynolds numbers
where the deposition decrease begins is well below the ranging from 400 to 21 000, the deposition rate (in fig/
critical temperature for most tests. Since the deposition (cm2*h) appeared relatively constant in the 225-325 °C
decrease occurs at differing temperatures at different flow (435-615 °F) wall temperature range (Marteney and
3120 Ind. Eng. Chem. Res., Vol. 32, No. 12, 1993

Table II. Autoxidation Deposition Peak (Averaged over


5-cm Tube Segment) Parameters for JP-7 as a Function of
Flow Rate*

parameters for flowrate, mL/min


autooxidation deposition peak 3.1 12.0 33.5
wall temperature, °C 390 441 523
fuel temperature, °C 294 266 225
residence time to peak, s 7 2 0.8
Reynolds number at peak 200 700 1500
peak deposition, ppm/cm2 0.0060 0.0037 0.0036
peak deposition, ng/(cm2-h) 1 2 5

Deposition values reported are for longest test reported, with


0

background (7 ng/cm2) subtracted.

Spadaccini, 1986). Expressed as ppm/cm2, the UTRC


deposition levels decrease with increasing flow rate. The
tests were not carried out to high enough wall temperatures
for the lowest and highest flow rates to determine if the Wall temperature, 'C
deposition decrease shifted with varying flow rate. Aside Figure 7. 12 mL/min deposition results for n-hexadecane and JP-7.
from the UTRC data, no published results for varying Air-sparged fuels, 69 atm.
flow rates were found which would confirm this shift in
the autooxidative peak location. However, recent work at
Southwest Research Institute has produced similar results
(Moses, 1992).
Focusing on the wall temperature alone is an obvious
oversimplification, since the deposition (as a chemical
process) is also a function of bulk fuel temperature and
residence time. Flow velocity may also be important for
deposit erosion. The importance of other parameters is
apparent in results obtained by Shell Development (Faith
et al., 1971), where two deposition peaks (autooxidative
and pyrolytic) are seen in a heated tube test where the
wall temperature is relatively constant at 650 °C (1200
°F), but the fuel temperature increases from 260 to 510
°C (500-950 °F). It seems likely also that deposition peaks
could form in tests with constant fuel and wall temper-
atures, where the peaks would arise from consumption
and/or exhaustion of reactive species. Nozzle fouling
simulations are usually performed with preheated fuel
flowing through hot metal tubes at relatively constant Figure 8. 3.1 mL/min deposition results for n-octane and JP-7.
temperature. Under these conditions, a deposition peak Air-sparged fuels, 69 atm.
may be absent (Chin and Lefebvre, 1992; Hazlett, 1991).
Another way of testing the supercritical solvency hy-
pothesis is to test fuels of similar structure but varying
critical temperature. This was done by testing n-octane
(Tc = 296 °C (564 °F)) and n-hexadecane (Tc = 447 °C
(836 °F)) under conditions similar to those for the JP-7
tests. The results are shown in Figures 7 and 8. Again,
it can be seen that the critical temperature of the fuel does
not appear to control the deposition behavior in these
tests. The deposition decrease occurs at essentially the
same temperature (both fuel and wall) for all of the
hydrocarbons, regardless of their critical temperature. The
deposition levels for the pure hydrocarbons are very similar
to that of JP-7.
The deposition rates of typical fuels are much higher
than that of JP-7 or pure hydrocarbons (Bradley et al.,
1974; Taylor, 1974; Frankenfeld and Taylor, 1980). The
thermal stability of an aircraft fuel can be assessed by the
temperature at which it fails the specification jet fuel
thermal stability test (the Jet Fuel Thermal Oxidation Figure 9. Deposition results for various fuels. Air-sparged fuels, 12
Tester (JFTOT), ASTM 3241). This is referred to as the mL/min, 69 atm.
breakpoint temperature. Thus, Taylor measured a dep-
osition of 0.4 ppm/cm2 for a JP-5 (274 °C breakpoint) Figures 4-8. In Taylor’s test, the total carbon deposit in
(Taylor, 1974) and UTRC measured a deposition of 0.04 the tube from JP-5 was 10 times that of JP-7 (Taylor,
ppm/cm2 for another JP-5 (breakpoint 270 °C) (TeVelde 1974). The results in the present test for a Jet A-l fuel
and Glickstein, 1983), as shown in Figure 1. These (333 °C breakpoint) and JP-7 are shown in Figure 9. The
deposition levels are an order of magnitude larger than deposition level is similar to that reported by Taylor for
the levels seen in JP-7 and the pure hydrocarbons in a JP-5 fuel at a similar flow rate (Figure 1). More extensive
Ind. Eng. Chem. Res., Vol. 32, No. 12, 1993 3121
Table III. Results of Gas Chromatographic Analysis of radicals is favored as the secondary C-H bond strength
n-Octane and n-Hexadecane after Thermal Stressing is 3 kcal/mol less than the primary C-H bond. There are
GC area % three possible sec-CgHn radicals that can be formed by
n-octane test n-hexadecane test
H-atom abstraction from n-octane. These three sec
species
radicals can recombine to yield six possible Cs dimers.
propane and propylene 0.125
The fact that we found six dimer species as products further
1-butene and n-butane 0.241/0.274 0.052
1-pentene 0.352 0.056 confirms their identity.
n-pentane 0.328 0.089 The major products found, the homologous series of
1-hexene 0.448 0.109 alkanes and alkenes, are similar to those observed for GC
n-hexane 0.048 0.101 analyses in the low-temperature (160-180 °C) reaction of
1-heptene and n-heptane 0.188 0.197/0.106 n-dodecane (Zabarnick and Striebich, 1992), and in the
1-octene 0.182
n-octane 0.119 pyrolytic cracking of alkanes at high temperature. It is
n-octane and impurities 97.65 thought that these alkanes and alkenes are products of
1-nonene 0.178 the decomposition of alkyl hydroperoxides (Reddy et al.,
n-nonane 0.125 1988) at temperatures below 480 °C (900 °F) and from
l-decene 0.180 pyrolysis reactions above 480 °C (Hazlett, 1977). The alkyl
n-decane 0.129
1-undecene 0.207
hydroperoxides are produced from the autooxidation chain
n-undecane 0.130 mechanism, where the reactant alkane, RH, is converted
1-dodecene 0.196 to the alkyl hydroperoxide, ROOH:
n-dodecane 0.126
1-tridecene and n-tridecane 0.314
R + 02 —* R02
1-tetradecene 0.150
n-tetradecane 0.148 R02 + RH -> ROOH + R
1-pentadecene 0.107 The R radicals are initially produced by some poorly
n-pentadecane 0.066 understood initiation mechanism that is thought to involve
Cs dimers 0.275
n-hexadecane and impurities 95.99 oxygen and surface catalysis. These alkyl hydroperoxides
other unidentified products 0.071 0.94 are known to be produced in air-saturated hydrocarbons
at temperatures as low as «=95 °C (200 °F). The alkyl
tests with various fuels and pure hydrocarbons are hydroperoxides decompose by the following path:
underway (Edwards and Liberio, 1994). ROOH — RO + OH
RO RCHO + R or RRCO + R

Chemical Analysis of Products


RO + RH -*• ROH + R
Gas chromatographic (GC) analyses were performed on R + RH -* alkane + R
samples of stressed n-octane and n-hexadecane for de-
termination of the reaction products. The products were R -* alkene + R
identified by a mass spectrometer detector interfaced to Thus, the alkyl hydroperoxides decompose to alkoxy
the GC and quantified using a flame ionization detector radicals and an OH radical. This alkoxy radical can
(FID). The n-octane sample was subjected to maximum decompose unimolecularly to aldehydes or ketones and
fuel and wall temperatures of 480 and 540 °C (900 and alkyl radicals, or abstract an H-atom from a fuel molecule
1000 °F), respectively (Figure 8). The n-hexadecane was to form an alcohol and an alkyl radical. Alkanes and
subjected to maximum fuel and wall temperatures of 470 alkenes are produced from H-atom abstraction from the
and 620 °C (875 and 1150 °F), respectively (Figure 7). The fuel and unimolecular decomposition, respectively. An-
product yields were estimated by assuming that the FID other possible pathway for the formation of alkenes is the
sensitivity to individual compounds is proportional to the intramolecular isomerization of alkylperoxy radicals to
number of carbon atoms present in the molecule. There- hydroperoxyalkyl radicals (QOOH), which can then de-
fore at small conversions of the parent alkane to products, compose to yield an alkene and an HO2 radical (Fish, 1970):
the integrated areas of the GC peaks can be related to the
fraction of the carbon atoms in the original alkane that R02 -* QOOH —*• alkene + H02
react. The products observed for these two experiments The alkyl hydroperoxide mechanism also predicts the
are shown in Table III. The data in the table indicate the formation of alcohols, aldehydes, and ketones, none of
3-4 % of the carbon of the reactant alkane has reacted and which was found in the GC analysis of the whole stressed
the primary products observed are a series of alkanes and fuel. At this time it is not clear why oxidized products
alkenes, with shorter carbon backbones than the parent were not found, although oxygenated products with the
alkane. For n-octane, a series of six peaks that appear to same carbon chain length as the original alkane would
be dimers of CgHn radicals, that is isomers of Ci6H34, are likely be masked by the very large alkane peak in the
also found. The mass spectra of these peaks do not yield analysis. Separation and concentration of the low con-
the molecular ion; identification as Cj6 species was obtained centration oxidized products using solid-phase extraction
by comparison of GC retention times with other alkanes, with subsequent GC/MS analysis of the extract (Reddy
comparison with mass spectra of other large alkane isomers, et al., 1988; Schulz, 1992) was successfully used to
the prominent C3Hi7 mass spectral peak, and the number qualitatively identify the oxidized products from the
of isomers found as discussed below. For n-hexadecane hexadecane. The oxidized products included the aldehyde,
the corresponding C32H66 species was not found, although alcohol, and ketone derivatives of the n-paraffins listed in
this could be due to these species not eluting from the Table III, such as the n-decane derivatives decanal,
chromatographic column due to their expected very high 2-decanone, and 1-decanol. The elution region near the
boiling points under the chromatographic conditions used oxidized hexadecane products was crowded with C15-C17
for these analyses. primary and secondary alcohols, ketones, and aldehydes,
The C16H34 isomers result from the recombination of as well as products identified as substituted dihydrofura-
sec-CsHi7 radicals. The formation of secondary alkyl nones.
3122 Ind. Eng. Chem. Res., Vol. 32, No. 12,1993

The chemical processes that result in deposit formation Edwards, T. “USAF Supercritical Hydrocarbon Fuels Interests”;
AIAA Paper 93-0807; AIAA: Washington, DC, January 1993.
are poorly understood. The absence of oxygen greatly
reduces deposit formation (Taylor, 1974; Bradley et al., Edwards, T.; Anderson, S. D. “Results of High Temperature JP-7
Cracking Assessment”; AIAA Paper 93-0806, AIAA: Washington,
1974; Hazlett, 1991). Hazlett (1977,1980) proposed alkyl DC, January 1993.
hydroperoxides as precursors to deposits due to the Edwards, T.; Liberio, P. D. The Thermal-Oxidative Stability of Fuels
observation that hydroperoxide depletion appeared to at 480 °C (900 °F). Paper be presented at Symposium on Distillate
correlate with tube deposit rating in a JFTOT. More Fuel Auto-Oxidation Chemistry, ACS National Meeting, San
recently, evidence has appeared demonstrating an inverse Diego, CA, March 1994.
correlation between the ability of a fuel to form peroxides Faith, L. E.; Ackerman, G. H.; Henderson, H. T. “Heat Sink
Capabilities of Jet A Fuel: Heat Transfer and Coking Studies”;
with its deposit-forming tendency (Heneghan and Zabar- NASA CR-72951; NASA Lewis Research Center: Cleveland, OH,
nick, 1992). Efforts are underway to incorporate oxygen July 1971.
consumption and solid deposition reactions into compu- Fish, A. Rearrangement and Cyclization Reactions of Organic Peroxy
tational fluid dynamic models of fuel system components Radicals. In Organic Peroxides; Swern, D., Ed.; Wiley-Inter-
(Katta and Roquemore, 1993). science: New York, 1970; Vol. I, pp 141-198.
Frankenfeld, J. W.; Taylor, W. F. Deposit Formation from Deoxy-
genated Hydrocarbons. 4. Studies in Pure Compound Systems.
Summary Ind. Eng. Chem. Prod. Res. Dev. 1980,19, 65-70.
Hazlett, R. N. Fuel Radical Reactions Related to Fuel Research. In
The results presented in this paper indicate that surface Frontiers of Free Radical Chemistry; Pryor, W. A., Ed.; Academic
deposition from supercritical hydrocarbon fuels is con- Press: New York, 1980; pp 195-223.
trolled by chemical processes in the fuel, rather than by Hazlett, R. N. Thermal Oxidation Stability of Aviation Turbine
supercritical solvent properties of the fuel. The primary Fuels; ASTM Monograph 1; American Society for Testing and
Materials: Philadelphia, 1991.
deposition property of interest was the decrease in
Hazlett, R. N.; Hall, J. M.; Matson, M. Reactions of Aerated
deposition usually seen at temperatures above approxi- n-Dodecane Liquid Flowing Over Heated Metal Tubes. Ind. Eng.
mately 370 °C (700 °F). The deposition decrease was Chem. Prod. Res. Dev. 1977,16, 171-177.
shown to be independent of fuel critical properties in two Heneghan, S. P.; Zabamick, S. Oxidation of Jet Fuels and the
ways. First, it was demonstrated that the deposition Formation of Deposits. Fuel 1993, in press.
decrease seen at high temperatures at a given flow rate Katta, V. R.; Roquemore, W. M. A Numerical Method for Simulating
could be shifted significantly in temperature by varying the Fluid-Dynamic and Heat Transfer Changes in a Jet Engine
the fuel flow rate. Thus, the deposition decrease is not Injector Feed Arm Due to Fouling. J. Thermophys. Heat Transfer
1993, in press.
solely a temperature effect, as it would be if supercritical Martel, C. R. Air Force Aviation Fuel Thermal Oxidation Stability
fuel properties were controlling the deposition. Second, R&D. In Jet Fuel Thermal Stability; NASA TM 79231; Taylor,
tests with fuels and pure hydrocarbons of widely varying W. F., Ed.; NASA Lewis Research Center: Cleveland, OH,
critical temperatures yielded very similar deposition- November 1978; pp 41-52.
temperature behavior, again demonstrating the high- Martel, C. R. Molecular Weight and Average Composition of JP-4,
JP-5, JP-8, and Jet A. Unpublished Results from Air Force Fuel
temperature deposition decrease is not primarily related Surveys. Unpublished memo, Air Force Aero Propulsion Labo-
to fuel supercritical properties. ratory, July 15,1988.
Marteney, P. J.; Spadaccini, L. J. Thermal Decomposition of Aircraft
Acknowledgment Fuel. J. Eng. Gas Turbines Power (ASME Trans.) 1986,108,648.
Moses, C. Southwest Research Institute. Unpublished work on NASA
This work was supported by the Air Force Office of Grant NAG 3-1342, 1992.
Scientific Research (AFOSR). Julian Tishkoff was the Reddy, K. T.; Cemansky, N. P.; Cohen, R. S. Modified Reaction
AFOSR Program Manager. The authors wish to thank Mechanism of Aerated n-Dodecane Liquid Flowing over Heated
Metal Tubes. Energy Fuels 1988, 2, 205-213.
Bill Schulz for his separation and analysis of the hexa-
Schulz, W. D. Oxidation Products of a Surrogate JP-8 Fuel.
decane oxidation products.
Prepr.—Am. Chem. Soc., Div. Pet. Chem. 1992, 37 (2), 383-392.
Taylor, W. F. Deposit Formation from Deoxygenated Hydrocarbons.
Literature Cited 1. General Features. Ind. Eng. Chem. Prod. Res. Dev. 1974,13,
133-138.
Bradley, R.; Bankhead, R.; Bucher, W. “High Temperature Hydro- TeVelde, J. A.; Glickstein, M. R. “Heat Transfer and Thermal
carbon Fuels Research in an Advanced Aircraft Fuel System Stability of Alternative Aircraft Fuels, Vol. I”; NAPC Report
Simulator on Fuel AFFB-14-70”; AFAPL-TR-73-95; Air Force NAPC-PE-87C, November 1983, AD A137 404.
Aero Propulsion Laboratory: Wright-Patterson AFB, OH, April Zabarnick, S.; Striebich, R. Unpublished results, 1992.
1974.
Chin, J. S.; Lefebvre, A. H. Experimental Techniques for the Received for review May 10,1993
Assessment of Fuel Thermal Stability. J. Propulsion 1992,8 (6), Revised manuscript received September 7,1993
1152-1156. Accepted September 20,1993*
Edwards, T. “Deposition During Vaporization of Jet Fuel in a Heated
Tube”; AIAA Paper 92-0687; AIAA: Washington, DC, January
• Abstract published in Advance ACS Abstracts, November
1992. 1, 1993.

You might also like