Sie sind auf Seite 1von 12

Proceedings of the 2012 9th International Pipeline Conference IPC2012 September 24-28, 2012, Calgary, Alberta, Canada

IPC2012-90535
LARGE-SCALE HYDROGEN JET FLAME RADIANT FRACTION MEASUREMENTS AND MODELING
Isaac W. Ekoto Sandia National Laboratories Livermore CA, USA William G. Houf Sandia National Laboratories Livermore CA, USA Adam J. Ruggles Sandia National Laboratories Livermore CA, USA

Leonard W. Creitz Air Products and Chemicals Inc. Allentown, PA, USA

Jimmy X. Li Air Products and Chemicals Inc. Allentown, PA, USA

ABSTRACT Analytic methods used to establish thermal radiation hazard safety boundaries from ignited hydrogen plumes are based on models previously developed for hydrocarbon jet fires. Radiative heat flux measurements of small- and mediumscale hydrogen jet flames (i.e., visible flame lengths < 10 m) compare favorably to theoretical calculations provided corrections are applied to correct for the product species thermal emittance and the optical flame thickness. Recently, Air Products and Chemicals Inc. commissioned flame radiation measurements from two larger-scale hydrogen jet flames to determine the applicability of current modeling approaches to these larger flames. The horizontally orientated releases were from 20.9 and 50.8 mm ID pipes with a nominal 60 barg source pressure and respective mass flow rates of 1.0 and 7.4 kg/s. Care was taken to ensure no particles were entrained into the flame, either from the internal piping or from the ground below. Radiometers were used to measure radiative heat fluxes at discrete points along the jet flame radial axis. The estimated radiant fraction, defined as the radiative energy escaping relative to chemical energy released, exceeded correlation predictions for both flames. To determine why the deviation existed, an analysis of the data and experimental conditions was performed by Sandia National Laboratories Hydrogen Safety, Codes and Standards program. Since the releases were choked at the exit, a pseudo source nozzle model was needed to compute flame lengths and residence times, and the results were found to be sensitive to the formulation used. Furthermore, it was thought that ground surface reflection from the concrete pad and steel plates may have contributed to the increased recorded heat flux values. To quantify this impact, a weighted multi source flame radiation model was modified to

include the influence of planar surface radiation. Model results were compared to lab-scale flames with a steel plate located close to and parallel with the release path. Relative to the flame without a plate, recorded heat flux values were found to increase by up to 50% for certain configurations, and the modified radiation model predicted these heat fluxes to within 10% provided a realistic steel reflectance value (0.8) was used. When the plate was heavily and uniformly oxidized, however, the reflectance was sharply attenuated. Model results that used the surface reflectance correction for the larger-scale flames produced good agreement with the heat flux data from the smaller of the two flames if an estimated reflectance of 0.5 was used, but was unable to fully explain the under predicted heat flux values for the larger flame. NOMENCLATURE af = Planks mean absorption coefficient [=0.23 for H2] Af = Flame surface area [m2] C* = Non-dimensional radiant power D = Distance [m] d, d* = Nozzle and mass weighted effective diameter [m] g = Gravitational acceleration [=9.81 m/s2] Frf = Flame Froude number L* = Non-dimensional flame length Lf , Lvis = Computed and measured visible flame length [m] = Mass flow rate [kg/s] MF = Stoichiometric fuel/air molecular weight N = Number of source points p = Pressure [kPa] Q, q = Heat release [kW], Radiative heat flux [kW/m2] R = Radius [m] RH = Relative humidity [%] Ru = Universal gas constant [=8314.5 J/(Kkmole)]

Copyright 2012 by ASME

S T Tad tf u VF w Wf ys Hc

= Surface emissive power [kW/m2] = Temperature [K] = Hydrogen adiabatic flame temperature [=2390 K] = Residence time [ms] = velocity [m/s] = View factor = Weight parameter = Flame width [m] = Stoichiometric mass fraction = Heat of combustion [=119 MJ/kg for H2] = Angle between surface normal and line of sight = Reflectance = Wind direction = Density [kg/m3] = Radiant fraction

the different modeling methodologies is given by Hankinson and Lowesmith et al. [12]. These models require knowledge of relevant release conditions (e.g., exit velocity, nozzle

Sub and superscripts 0 = Stagnation condition amb = Ambient condition clip = Clipped surface condition eff = Effective pseudo source condition j = Jet Exit condition inf = Infinite surface condition meas = Measured quantity obs = Observer quantity ray = Ray trace vector quantity refl = Surface reflector quantity INTRODUCTION Renewed interest in hydrogen as an energy carrier for portable power generation and vehicle transport has necessitated the development of new hydrogen storage, transport, and delivery infrastructure; which themselves introduce new operational considerations that must be addressed prior to widespread deployment. Of particular concern according to LaChance et al. [1] is the large-scale unintended release and subsequent ignition of compressed hydrogen into the ambient. For personnel in the vicinity of the hydrogen jet fire, radiant heat flux exposures can result in potentially lethal burns, while elevated air temperatures can lead to severe respiratory damage. To ensure suitable safety measures for personnel and equipment are in place, probabilistic risk assessments (PRA) derived from failure modes and effects analyses [2] are commonly used. The PRA approach, however, strongly depends on validated consequence modeling of relevant flame behavior over a broad range of potential release scenarios. Detailed modeling of flame behavior via computational fluid dynamic (CFD) simulation provides useful information regarding the interplay between flow dynamics and combustion chemistry [3-5]. However, the significant computational time required along with the necessary access to and familiarization with suitable simulation packages make CFD prohibitive for practical safety applications. Instead, reduced order jet-flame radiation models developed for hydrocarbon flames and based on empirical observation are typically used [6-12]; a review of

Figure 1. Image of the pipe release setup at the GL Noble Denton Spadeadam Test Site (top), still images from jet flames 1 and 2 (middle), and a schematic of the storage reservoir and release setup (bottom).

Copyright 2012 by ASME

diameter/shape, mass flow rate, gas type) to determine the flammable envelope along with an estimate for the amount of flame energy that is converted into escaping radiant energy, which is defined here as the radiant fraction. Hydrocarbon jet fire thermal radiation mostly results from the combination of species specific thermal emittance from hot combustion products (i.e., H2O and CO2) at discrete wavelengths and broadband blackbody emission from heated soot particulates. Both emission types are generally in the infrared portion of the light spectrum, although a small amount of visible and ultraviolet emission also occurs. Since hydrogen jet fire combustion produces no CO2 or soot in the product stream, radiant fractions are thought to be lower. Nonetheless, Schefer et al. [13] reported that like hydrocarbon flames, measured radiant fractions from laboratory scale hydrogen jet flames exhibit a logarithmic dependence on flame residence time [14]. Molina et al. [15] was further able to develop a unified expression for non-sooty hydrocarbon and hydrogen flame radiant fractions as a function of flame residence time, provided corrections for the adiabatic flame temperature and the product species Planks mean absorption coefficient were applied. However the modeling approach has only been verified for small and medium sized hydrogen flames (visible flame length < 10 m), and the applicability to larger scale flame types remains uncertain. In 2008, Air Products and Chemicals Inc. commissioned experimental measurements from two large-scale, horizontally orientated, hydrogen flames to validate current modeling approaches for these larger flames. An initial analysis of the data suggested that the radiant fraction derived from distributed heat flux sensor measurements was up to 40% greater than model predictions. Sandia National Laboratories Hydrogen Safety, Codes and Standards research group has independently analyzed these results and the experimental setup used to acquire the data to determine the reason for the discrepancy. The present paper discusses measurement details and the potential impact of reflected emittance on measured heat flux. From the analysis, enhancements to current prediction methodologies have been proposed. LARGE-SCALE FLAME EXPERIMENT DESCRIPTION Two large-scale jet fire experiments were conducted at the GL Noble Denton Spadeadam Test Site in North Cumbria, UK. For each experiment, compressed hydrogen gas from a nominal 60 barg stagnation pressure was released through a horizontally orientated 1 m long stretch of pipe with respective internal diameters of either 20.9 mm (Jet 1) or 52.5 mm (Jet 2), and located 3.25 m above the ground. An image of the release setup, along with a schematic of the delivery system is shown in Figure 1. Directly in front of the release location was an approximate 25 m by 15 m concrete pad that was used to prevent the entrainment of surface dirt particles. The pad was further covered with steel sheeting directly beneath the release path to protect it from spallation. Since the gas system had previously been used with natural gas, 3 consecutive hydrogen purges from 0 to 7 barg were

performed to remove residual natural gas from the storage and delivery system. Separate purges of compressed hydrogen and nitrogen were also used to clear debris and corrosion products from the internal portion of pipework connecting the reservoir to the nozzle. A gas sample taken at the end of the purge cycles was sent to an external laboratory for analysis; results indicated the reservoir methane concentration was less than 0.02%. Pipework connecting the gas storage reservoir to the jet release point was fitted with a flat plate orifice and remotely operated flow control valve, shown schematically in Figure 1. Mass flow rates for each release were calculated from upstream temperature measurements and the pressure drop across the orifice plate. Installation of the orifice plate and calculation of the gas mass flow rates were performed according to ISO 5167 parts 1 and 2 [16]. The pressure drop was measured by a Druck STX 2100 differential pressure transducer with a 0 2.0 bar range and 0.2% full-scale accuracy. Static temperature was measured by a type T thermocouple with outputs linearized via a Pretop 5331B temperature transmitter with a programmed range of 100C and a full-scale accuracy of 0.05%.

Figure 2. Schematic of the radiometer and camera placement relative to the jet release point. Coordinates for all radiometers are given in Table 2 Table 1. Boundary and ambient conditions for each largescale jet flame. Note that wind directions are where the wind is coming from while the release direction is the direction of the release path.
Jet dj p0 T0 [mm] [kg/s] [barg] [K] 1.0 7.4 RH [%] Tamb pamb uwind [K] [mbar] [m/s] 280 280 1022 1011 2.84 0.83

wind
[] 68.5 34.0

Lvis [m] 17.4 48.5

1 20.9 2 52.5

59.8 308.7 94.3 62.1 287.8 94.5

Static pressure and temperature were measured at three locations in and around the release pipe as shown schematically in Figure 1. Because of the large internal area ratio of the upstream 6 pipe section relative to the release pipe, the static pressure and temperature measurements at this location were

Copyright 2012 by ASME

assumed to be very near stagnation conditions. Static pressure was measured via Druck PTX-1400 pressure transducers with a 0100 barg range and full-scale accuracy of 0.15%, while static temperature was measured using the same thermocouple types used in the in the upstream pipe work for the mass flow rate calculations. Details of the mass flow rate and stagnation conditions for each release are summarized in Table 1. Incident thermal radiation was measured at 13 locations by wide-angle (150 field of view) Medtherm radiometers mounted on tripods around the test area. Location details of each sensor are schematically illustrated in Figure 2, and given in Table 2 for both jets. Each radiometer was fitted with a calcium fluoride window that transmitted light between 0.3 to 11.5 m, had a 1 second response time, and a full-scale accuracy of 5%. Prior to the measurements, each sensor was calibrated in a blackbody furnace. Four radiometers were placed perpendicular to the predicted flame center, at different locations in the positive z-axis (see Figure 2). An additional sensor was placed on the opposite side to check for flame symmetry. Four radiometers, two on each side, were placed 5 m downstream from the predicted flame center. Two additional radiometers were placed far downstream from the release axis, and orientated 45 relative to the release axis. All other radiometers along the flame side were orientated normal to the release axis. Finally a sensor was placed both directly downstream and upstream from the flame, and orientated along the release axis. It should be noted that due to the close proximity of sensor RAD13 to the flame, it was possible that the visible flammable extent extended beyond the view area. However, because of the large sensor field of view, this possibility was deemed unlikely. Table 2. Large-scale jet flame radiometer position relative to the release exit. Sensor RAD01 RAD02 RAD03 RAD04 RAD05 RAD06 RAD07 RAD08 RAD09 RAD10 RAD11 RAD12 RAD13 x [m] 11.0 11.0 11.0 11.0 16.0 16.0 21.6 26.0 21.6 16.0 16.0 11.0 -1.0 Jet 1 y [m] -2.4 -1.7 -1.9 -0.3 -2.4 -1.8 -1.8 -1.5 -1.2 -1.3 -0.5 -1.8 -1.0 z [m] 7.5 12.5 15.0 17.5 7.5 12.5 10.6 0.0 -10.6 -7.5 -12.5 -7.5 0.0 x [m] 23.0 23.0 23.0 23.0 28.0 28.0 40.7 48.0 40.7 28.0 28.0 23.0 -1.0 Jet 2 y [m] -2.4 -1.7 -1.9 -0.3 -2.4 -1.8 -1.8 -1.5 -1.2 -1.3 -0.5 -1.8 -1.0 z [m] 15.0 20.0 30.0 40.0 15.0 30.0 17.7 0.0 -17.7 -15.0 -30.0 -15.0 0.0

down-stream field of views. Relative positions of all cameras are shown in Figure 2. Wind speed/direction, ambient temperature, and relative humidity were measured at a weather tower located ~111 m upstream from the release point, as indicated in Figure 2. An additional relative humidity sensor was collocated with radiometer RAD01 to ensure the measurements were consistent. Ambient pressure was reported from a nearby weather station located at Carlisle, Cumbria, UK. Ambient conditions for each release are reported in Table 1. Note that the jet release was oriented 67 relative to true north, and wind directions reported in Table 1 are likewise relative to true north. Prior to each experiment, the storage reservoir was charged with 90 barg of compressed hydrogen, supplied by a series of Air Products tankers. Experiments were initiated by opening of actuated valves to allow the gas to flow from the storage reservoir to the release pipe. Flame ignition was remotely initiated by an incendiary device positioned close to the release pipe. All recorded data and video were synchronized to the incendiary device initiation. Because of the turbulent nature of the studied flames and the variability of the atmospheric conditions, measurements were acquired over discrete averaging periods to ensure representative mean values were acquired. For Jet 1, the averaging period was between 50 and 90 seconds after spark ignition, when the stagnation pressure was nominally steady at around 60 barg. Because of the limited storage capacity of the storage reservoir and the high flow rates achieved for Jet 2, it was not possible to maintain a constant 60 barg back pressure. Instead, the flow control valve was fully opened and the stagnation pressure was allowed to monotonically decrease. The decay rate in the back pressure was sufficiently slow such that the release could be treated as quasi-steady over a small averaging period. For the present measurements, the averaging period select was between 10 and 20 seconds after ignition. For this period, the stagnation pressure decreased from 63.1 to 56.6 barg with an average pressure of 59.8 barg. LAB-SCALE FLAME EXPERIMENT DESCRIPTION To explore the potential contribution of reflected radiation on recorded heat flux values, a vertically orientated hydrogen jet flame with a similar 60:1 pressure ratio used for the largescale experiments was examined in Sandia/CAs Turbulent Combustion Laboratory burner facility [17]. The jet issued from a high-pressure stagnation chamber and through a 1 mm diameter nozzle at the chamber outlet with a machine tapered profile [18] that was manufactured from a Swagelok one inch VCO fitting blank (SS-16-VCO-1-16). The estimated visual flame length was 1.0 m. Chamber temperature and pressure were respectively monitored via a type K thermocouple and TESCOM series 100 pressure transducer. Dynamic feedback was used to maintain a constant pressure ratio (5.964 MPa) and temperature (295.2 K). Data acquisition and system control were handled via a custom written LabView virtual instrument. Five Medtherm 64 series heat flux radiometers that were fitted with a zinc selenide optical window and had an

Standard video records of each flame were acquired at 3 locations around the test area. The flame envelope was measured using two thermal imaging cameras (FLIR Thermacam P65) positioned perpendicular to the cross- and

Copyright 2012 by ASME

Finally, one side of the plate was heavily and uniformly oxidized to assess the impact of oxidation on reflection. FLAME RADIATION MODELS The amount of heat flux that an observer receives can be expressed as the product of the view factor, VF, the surface emissive power of the flame, S. Eq. 1. where was the atmospheric transmissivity, which was expressed by simple Beer-Lambert expressions to account for the amount of CO2 and H2O in the view path [20]. Here, the view factor is defined as the portion of radiant emission viewable by an observer. Surface emissive power, or the amount of radiant energy emitted per unit area, is a function of the total flame heat release, Q, and the radiant fraction, , i.e., Eq. 2. where Af is the flame surface area, and flame heat release is the product of the mass flow rate, , and gas heat of combustion, Hc (= 119 MJ/kg for H2). A relation for is then obtained by combining equations 1 and 2: Eq. 3.

Figure 3. Illustration of lab-scale radiometer positions relative to the jet centerline and steel plate, along with photos of the oxidized and non-oxidized steel plates. An image of the jet flame is also included. acceptance angle of 150 were used to record the heat flux from the hydrogen flame. In the axial direction, sensors Rad1-Rad3 were placed at half the visible flame height (x = 0.5 m), while Rad4 was located 0.14 m directly downstream from Rad2. All sensors were located at half the visible flame length (r/Lvis = 0.5) from the flame centerline axis, with sensors Rad1 and Rad3 placed 20 from Rad2. Location details of each radiometer are shown schematically in Figure 3. Optics tables and potentially reflective surfaces were covered with black cloth to minimize any undesired reflections. To convert the recorded signal voltage into temperature a blackbody device (Mikron M330) was used. Each sensor was placed flush against a 2.54 cm diameter orifice at the device exit for a series of nine temperatures starting at 1050C decreasing in 100C intervals to 250C. For each temperature setting over 700 samples were collected. A temperature calibration curve was made by fitting a 4th order polynomial to the data. Radiometer temperature measurements were converted to radiant emittance through the use of a transmission function that accounted for the zinc selenide optical filter [19]. To assess the impact of reflectance a 610 mm 305 mm steel plate was mounted in a landscape orientation with the axial center aligned with the axial position of sensors Rad1Rad3 (x = 0.5 m) and at varying perpendicular distances away from the jet centerline. Starting at y = 200 mm, the plate was moved in the positive y direction in 50 mm increments. The plate was positioned such that the horizontal midpoint was coincident with the jet centerline (z = 0 mm), and subsequently was moved towards the group of radiometers such that the plate edge was coincident with the jet centerline (z = -305 mm).

Because the quantity VF is unknown, direct calculation of via the heat flux measurement is not possible without an assumed flame shape to determine total surface area along with the orientation of the flame relative to the observer. For the original data analysis, a single point flame radiation model from Sivathanu and Gore [8] was used, along with a nondimensional radiant power term, C*, that was used to account for flame shape at different downstream locations. Although originally validated against hydrocarbon flames, the model was later verified for small- and medium-scale hydrogen flames by Schefer et al. [13]. The ratio of flame surface area to the view factor and transmissivity from equation 3 was represented as: Eq. 4.
( )

where the radius, R, was the distance between the radiometer axial location and the flame centerline. Although C* depends on both the radial and axial distances of the observer relative to the flame, in the far field (r/Lvis > 0.5) the radial dependences largely disappears and C* can be expressed as: Eq. 5. ( | |)

Recently Hankinson and Lowesmith [12] proposed a weighted multi point approach to model radiation. Using this approach, C* would be: Eq. 6.

Here, Di is the distance between the observer and the point source i, while i is the angle between the observer normal and

Copyright 2012 by ASME

the vector between the observer and source points. The weighting parameter, wi, is defined by: Eq. 7. * where N was the total number of distributed points. For the present study a value of N = 80 was found to be suitable to ensure convergence of the results. The Hankinson and Lowesmith correlation has been found to produce far more reliable results in the near field of the flame radiation boundary relative to single point source models. It should be cautioned that since radiant fraction calculations are highly sensitive to the assumed flame shape and the original radiant fraction correlations (described below) were developed using the Sivathanu and Gore single point source model, the Hankinson and Lowesmith weighted multi point model may not properly match the developed correlation curve. However, the assumed flame shape is similar for both methods, and radiant fractions calculated from hydrocarbon jet flame heat flux measurements in the far field using both methods exhibit good agreement [12]. For completeness, computed radiant fractions using both methods are provided for most points in Table 3. A schematic of both the single and weighted multi point source models is shown in Figure 4. Note that equation 5 is not applicable for heat flux measurements from radiometers 8 and 13 located along the jet release axis. The applicability of equation 6 is questionable as well for these points, and thus these values are not shown. Furthermore, since radiometers 7 and 9 were orientated 45 relative to the flame axis, computed radiant fractions from these measurements were likewise excluded for the single point source method calculations. ( )+

To perform a predictive radiant fraction calculation, a correlation based on the flame residence time, tf, or the time spent by the fuel in the active flame zone has been developed by Molina et al. [15]. The formulation from Turns and Myhr [14] has been used where flame residence time was expressed as the ratio of fuel mass in the flammable zone to the mass flow rate: Eq. 8.

Here, ys, is the stoichiometric mass fraction, and f is the flame density (kg/m3) given by pambMF/RuTad (pamb is the ambient pressure, MF is the stoichiometric fuel/air molecular weight, Ru is the Universal Gas Constant, and Tad is the adiabatic flame temperature). Since the flame is assumed to be conical, Lf and Wf represent the computed visible flame length and flame width base respectively. The visible flame length was estimated from the length scale correlation by Kalghatgi [21]: Eq. 9.

and the computed flame width base was taken as Wf = 0.17Lf. Here the non-dimensional flame length, L*, is expressed as a function of the flame Froude number Eq. 10. where, Eq. 11.
( ) ( ) ( )

Figure 4. Schematic illustrating the single and weighted multi source models used by equations 5 and 6 respectively to compute the non-dimensional radiant power, C*.

From equation 9, it is apparent that the flame length scale, Lf, is proportional to the mass weighted nozzle diameter, d* ( ), where deff and eff are the effective jet exit diameter and density derived from application of an appropriate pseudo source model that accounts for complex jetexit shock structure. Several pseudo source models that incorporate various levels of complexity have been proposed [22-27] and applied to various compressed gaseous hydrocarbon releases. However, unlike hydrocarbon gases, pressurized hydrogen exhibits strong non-ideal behavior, and thus the formulating assumptions for each model is in doubt. The ability of 5 different pseudo source models to predict downstream characteristics of a hydrogen jet from a compressed storage vessel (10 bar storage pressure) was performed by Ruggles and Ekoto [17]. Each model was updated with a non-ideal equation of state (Abel-Noble) and the calculated d* was compared to the measured value. Models that excluded momentum conservation from the derivation and instead specified a velocity and temperature at the Mach disk [22, 24], over predicted d* by ~10%. On the other hand, models that included momentum conservation and had an assumed temperature at the Mach disk [23, 25], under predicted d* by

Copyright 2012 by ASME

~10%. Finally, the most complex model that included all conservation equations and accounted for the entropy changes across the shock over predicted d* by ~60%, although this was likely due to the unrealistic assumption that all fluid passed through the Mach disk. A comparison of the pseudo source models is discussed in the next section. Compared to hydrocarbon jet flames, radiant fractions with similar residence times are significantly lower for hydrogen jet flames. Molina et al. [15] attributed the difference to the value of the Plank mean absorption coefficient, af (=0.23 for hydrogen/air flames) in the product gases and accordingly developed the following unified radiant fraction correlation that was independent of fuel type: Eq. 12. (
)

LARGE-SCALE FLAME RESULTS AND DISCUSSION A still image from standard video recordings of the two large-scale jet flames is provided in Figure 1. Jet 2, with its larger release diameter and mass flow rate had a substantially larger flammable extent relative to Jet 1 (48.5 m vs. 17.4 m visible flame length). It was observed that this flame had a significant amount of visible flame luminosity. The bright yellow and orange colors were deemed to be the result of saturated lines in the visible emission bands, which typically are not observed for smaller-scale hydrogen flames [28]. It is also worth noting that there was a large amount of reflected visible radiation seen from the surface below the release. Table 3. Radiative heat flux measurements and radiant fractions calculated using both the single point (equation 5) and weighted multi point (equation 6) C* formulations. Jet 1 Sensor RAD01 RAD02 RAD03 RAD04 RAD05 RAD06 RAD07 RAD08 RAD09 RAD10 RAD11 RAD12 RAD13 qmeas [kW/m2] SPS 15.2 6.9 5.0 3.6 11.4 5.6 4.6 4.7 5.4 17.7 8.1 21.2 10.4 0.13 0.15 0.16 0.15 0.21 0.27 N/A N/A N/A. 0.31 0.38 0.17 N/A WMP 0.14 0.15 0.15 0.14 0.15 0.15 0.13 N/A 0.15 0.22 0.21 0.19 N/A Jet 2 qmeas [kW/m2] SPS 39.4 28.9 12.7 7.0 36.6 12.9 20.0 23.9 21.0 44.1 13.4 38.6 18.1 0.28 0.36 0.35 0.34 0.19 0.27 N/A N/A N/A 0.23 0.28 0.27 N/A WMP 0.24 0.27 0.23 0.22 0.20 0.23 0.18 N/A 0.18 0.24 0.24 0.23 N/A

from measurement points radially situated within half the flame length, while points 7 and 9 were excluded because they were not orientated normal to the flame axis); these values are summarized in Table 4. Average radiant fractions from the single point source method (equation 5) [8, 13] were 0.22 for Jet 1 and 0.32 for Jet 2. It should be noted, however, that a large spread in the data was observed. For the weighted multi source model (equation 6), the average radiant fractions were considerably lower at 0.16 and 0.22 for Jets 1 and 2 respectively. Furthermore, the variation from the mean values was within 20% for most points. Because of the more consistent results, the values from the weighted multi source model were deemed to be a closer representation of reality. Modeled radiant fraction calculations were performed using equations 8 through 11 along with a non-ideal gas formulation (Abel-Noble) of 4 of the pseudo source models briefly described in the previous section [22-24, 26]. These values are tabulated in Table 4 for both jets. Also included in Table 4 are the calculated flame lengths from equation 3. For Jet 1 with the smaller pipe diameter and mass flow rate, the computed radiant fraction was between 0.13 and 0.15, which slightly under predicted the measured value of 0.16. For Jet 2 with the larger flow rate and release diameter, the calculated radiant fraction was between 0.15 and 0.17 depending on the pseudo source model used, which resulted in an under prediction of the measured radiant fraction by ~30%. Calculated heat flux values based on the weighted multi source method from equations 6 and 7 along with the use of 4 different pseudo source models to determine radiant fractions are compared to measured values in Figure 5. The dashed line represents the point where model predictions and measured values are equivalent. Many of the near-field heat flux values from Jet 1 and nearly all the heat flux values from Jet 2 were under predicted, irrespective of the pseudo source model used. Table 4. Measured radiant fractions from the single (equation 5) and weighted multi point (equation 6) C* formulations compared to modeled radiant fractions from equation 12 with inputs from 4 notional nozzle models. Also included are computed flame lengths from equation 9. Jet 1 Sensor
Meas. [eqn. 5] Meas. [eqn. 6] Birch et al. (1984) Ewan & Moodie Birch et al. (1987) Harstad & Bellan

Jet 2 Lf q
m

0.22 0.15 0.14 0.13 0.14 0.15

Lf 49.2 49.9 44.6 52.5

20.5 21.0 18.6 23.5

3 0.32 2 9 0.22 8 0.16 1 2 0.15 7 . 0.16 3 6 0.17 1 2

Radiant fractions were computed from an average of the radiant fraction values in Table 3 for sensors 1-7 and 10-12 (note that points 1, 5, 10, and 12 were excluded from radiant fraction calculations that used equation 5 because they were

Due to the close proximity to the of the jet fire to the ground, it may be possible that reflected radiation artificially increased the recorded jet flame heat flux. As a justification for this hypothesis, a conceptual diagram of a radial cut plane

Copyright 2012 by ASME

relative to an observer at arbitrary height is shown in Figure 6. From this schematic, if a reflective surface is present on the ground, the observed flame radiation will be the sum of both the incident and reflected radiation. The amount of radiation that is reflected will depend on several factors including: the height of the observer above ground, the height of the flame above the ground, the extent of the reflective surface, the distance of the observer from the flame, and the reflectance of the surface material.

controlled laboratory experiments and a model representation of surface reflection are discussed in the next section.

Figure 6. Conceptual schematic of radiative heat flux addition via reflected surface radiation, where obs and refl represent the angular deviation between the observer and flame surface normals from the ray connecting each point. LAB-SCALE FLAME RESULTS AND DISCUSSION Given the presence of a reflected surface adjacent to a radiative emitter, there are three possible scenarios: 1) no reflected emission will be viewed by the observer, 2) all reflected emission will be viewed by the observer, or 3) a clipped portion of the reflected emission will be viewed by the observer. Reflective addition for scenario 1 does not need to be considered, while for scenario 2 it can be treated using the same C* formulation from equation 6, provided the centerline radial distance, R, and emitter distance, Di, are represented by the total magnitude of the line-of-sight vectors connecting the observer-to-reflector and reflector-to-emitter. However, point source models are insufficient to model scenario 3 since the physical flame shape is not considered. To get around this problem, the weighted multi source model was modified such that the source points were replaced by uniformly radiating spheres with radius, ri, given by: Eq. 13.

Figure 5. Comparison of heat flux measurements relative to calculations using 4 different notional nozzle models and the weighted multi source formulation for C* (equation 6). It was already pointed out that visible light was observed reflecting off the concrete/steel surface in Figure 1. However, since most radiative heat flux from hydrogen combustion is from the infrared portion of the light spectrum, information regarding the wavelength specific reflectance of the surface material is needed. Although to the best of the authors knowledge no information regarding the optical constants of carbon steel exists in the literature, Karlsson and Ribbing [29] noted that simple Ferritic and Martensitic stainless steels have similar reflectance values of ~0.8 to 0.9 in the far infrared (wavelengths > 2 m), despite varying Chromium content. Moreover, these values are in good agreement with near infrared values for pure iron from Yolken and Kruger [30]; thus it can safely be assumed that most carbon steels will likewise have a reflectance of about 0.8. McIlvaine et al. [31] reported near infrared reflectance values for different types of concrete to be about 0.3. It then seems reasonable that significant infrared flame radiation may have been from surface reflected irradiance. The amount reflected depended on the relative orientation and location of the radiometer, steel/concrete surfaces, and jet flame, along with overall surface condition. Ultimate quantification of the impact of surface reflection requires a validated model correction. Results of well-

Thus, the portion of reflective radiative emission received by the observer will then be proportional to the area ratio of the clipped view area, Aclip, and the area that would be observed if the reflective surface was infinitely wide, Ainf: The formulation for equation 6 can then be modified as: Eq. 14.

Here, the subscript, ray, denotes the magnitude of the ray trace vectors that connect the observer to the emitter surface, and is the surface reflectance. The clipped viewing area, Aclip, is determined by the integration of the visible radiating sphere surface area below the planar intersection of ray trace vector from the edge of the reflective surface. Note that the term, cosi,refl, was added to account for the deviation of the visible reflective surface unit normal from the reflection ray trace vector that connects the visible surface area to the observer. Lab-scale experiments discussed earlier were performed to verify the validity of the updated model. The top graph in Figure 7 shows a comparison the radiative heat flux from all sensors for all conditions, while all values are summarized in

Copyright 2012 by ASME

~5%. When plate with a 20 cm radial distance was offset in the negative z direction (see Figure 3) such that the edge of the plate furthest from the radiometers was aligned with the flame centerline, a dramatic radiative heat flux increase (up to 50%) was observed for all sensors. From the arguments given earlier, the heat flux increase indicates that the clipped visible flame area was approaching the visible area for an infinitely wide reflective surface. In contrast the previous results, Rad2 exhibited the highest heat flux increase. Finally, if the unoxidized plate for the initial configuration was replaced by the heavily oxidized steel plate, the change in relative heat flux relative to the measurements without a reflective surface was negligible. Table 5. Radiative flux data (kW/m2) for the lab-scale jet flame both with and without adjacent steel plates. Sensor (plate distance) No plate Oxidized Steel (20 cm) Steel Plate (20 cm) Steel Plate (30 cm) Offset Steel Plate (20 cm) 1 0.757 0.751 0.773 0.780 1.140 2 0.730 0.744 0.861 0.798 0.982 3 0.745 0.766 0.924 0.850 0.940 4 0.789 0.798 0.903 0.845 1.008

Figure 7. Top: Lab-scale vertical flame radiative heat flux measurements both with and without adjacent steel plates. Bottom: Model predictions relative to measurements, where a constant reflectance of 0.8 was assumed for all cases. Table 5. Heat flux values from sensors Rad1 to Rad3 were approximately equal, while the recorded heat flux from sensor Rad4 was somewhat larger due to the fact that it was more closely positioned to the thickest portion of the flame. When the un-oxidized steel plate was introduced 20 cm from the jet, sensor Rad1 had a negligible increase in heat flux, as most of the reflected heat flux likely bypassed the viewing area. However, heat flux values from sensors Rad2 through Rad4 exhibited a ~10% increase. When the radial distance of the plate was moved 30 cm from the jet centerline the heat flux increase observed for sensors Rad2 through Rad4 was lower at

Experimental results were compared modeled heat flux values from the updated model with the inclusion of surface reflection correction. In this case, the reflectance value was assumed to be 0.8 for all conditions, including for the unoxidized steel plate. The notional nozzle from Birch et al. [23] was used since this model produced radiant fraction (0.0104) and flame length (0.98 m) values closely matched experimentally observed values for the flame without a reflective surface (0.0098 and 1.0 m respectively). Model results relative to heat flux measurements are compared for all sensors at all conditions in the bottom graph of Figure 7. Despite the broad range of reflective surface orientations relative to the flame and heat flux sensor, the model values were within 10% for all conditions and at all sensors. The lone exception was for the oxidized plate, where the model over predicted the heat flux at sensors Rad2 to Rad4 due to the failure to account for diffusion by surface oxidant. Finally, model results for the large scale flames were updated to include the reflectance correction. A reasonable estimate for concrete reflectance was 0.3, while lab measurements indicated steel reflectance could be 0.8 for specular surfaces and much lower due if significant surface oxidation was present. Since the details of the steel plate surface conditions are unknown and the orientation of the flame centerline would have shifted from the release axis due to wind and buoyancy effects, it is impossible to properly assign a surface reflectance value or a relative contribution to the steel or concrete surfaces. Instead a reflectance value of 0.5 was assumed for a reflective surface with a 7.5 m radial extent, and the results are presented against the measured values in Figure 8 with the same 4 pseudo source models used to generate Figure 5. It should be noted that the reflective surface was treated as an infinite reflector in the axial direction, since

Copyright 2012 by ASME

the axial position of the flame relative to the concrete and steel plate surface was also unknown. This assumption should only have a minor impact, however, since the reflective surface covered the flame area with most of the radiative emission. For Jet 1 with the smaller release diameter and lower flow rate, the heat flux values are represented by the open symbols in Figure 8. For this condition, far-field measurements relative to Figure 5 were largely unchanged since the visible area ratio from equation 14 was close to or equal to zero for these points. Near-field radiometers were mounted along the edge of the reflective surface; hence all surface reflection was directed towards the observer. With the inclusion of surface reflection and the use of the assumed 0.5 reflectance, model predictions jumped by up to 50% for the near field locations and were found to be in line with the measured values. For Jet 2, the impact of surface reflection was less significant because of the larger separation distance between the edge of the reflective surface and the radial location of the radiometers. For this flame, the maximum reflective addition was ~25%, which occurred for the near-field sensors located in the negative z direction (Figure 2). Even with the inclusion of reflective addition, model predicted heat flux values were still well below the measurements. These results suggest reflective addition alone is not sufficient to fully explain the deviation between measurement and the model for the larger flame.

Figure 8. Comparison of heat flux measurements with reflection corrections applied. CONCLUSIONS Reduced order flame radiation consequence models are quick and reliable methods that can be used in conjunction with probabilistic risk assessment methods to establish harm safety distances from ignited flames. Since most available flame radiation models are specific to hydrocarbon flames, modifications are needed for hydrogen jet flames have only been verified for small and medium scale flames. To address

this gap for larger scale flames, Air Products and Chemicals Inc. commissioned two experiments from horizontally orientated pipes with exit diameters of 20.9 and 50.8 mm. each release was from a nominal 60 barg source pressure that produced a choked flow release and had respective mass flow rates of 1.0 and 7.4 kg/s. Care was taken to ensure no surface particles entrained into the flow. Radiant fractions were calculated using both a conventional single point and an updated weighted multi source models. More consistent radiant fraction calculations were observed with the weighted multi source model, and accordingly this model was the formulation primarily used for data reduction. Based on measurements from distributed radiometers flame radiation heat flux values exceeded model correlations by up to 30%. From an analysis of the data and experimental setup, two potential factors were identified. The first was the pseudo source model used to account for complex jet exit shock structure from the choked exit releases. These source models were used to solve for radiation model inputs (e.g., flame length and residence time). Available pseudo source models were updated with an Abel-Noble non-ideal gas equation of state. While calculated radiant fractions based on each model were found to be within ~10% of each other for the present flames, the model by Birch et al. [23] was found to be the most consistent predictor of visible flame length. The second identified factor was ground surface reflection from a concrete pad and steel plate below the release path. To quantify this impact, the weighted multi source flame radiation model was modified to include the influence of planar surface radiation, which was subsequently applied to well-controlled laboratory flame heat flux measurements with a steel plate located close to and parallel with the release path. Recorded heat flux values were found to increase by up to 50% for certain configurations, and these increases were well predicted by a surface reflection heat flux model provided a steel reflectance value of 0.8 was used. If a heavily oxidized steel plate was used in lieu of the clean un-oxidized steel, the reflectance addition became negligible. When applied to the larger scale flames, the surface reflectance model with an assumed 0.5 reflectance was able to explain near-field heat flux model deviations from the measurements for the smaller of the two large-scale flames. However, the modeled heat flux was still under estimated for the larger flame, which suggests that either the correlation between radiant fractions and flame residence time needs to be updated, or unidentified physical effects for very large-scale flames resulted in a more substantial heat flux increase above what would be expected otherwise. Additional large-scale measurements are needed to confirm these results. The current analysis suggests that for large-scale jet fire measurements, when possible, the flame should be orientated vertically to minimize the potential of ground surface radiation. Furthermore, reflected radiation can significantly increase the radiative heat load for a system with large reflective areas, and should accordingly be accounted for by system designers.

10

Copyright 2012 by ASME

ACKNOWLEDGMENTS This research was supported by the United States Department of Energy, Office of Energy Efficiency and Renewable Energy, Fuel Cell Technologies Program, under the Safety, Codes, and Standards subprogram element managed by Antonio Ruiz. Sandia is operated by the Sandia Corporation, a Lockheed Martin Company, for the U.S. DOE under contract No. DE-AC04-94-AL8500. The authors gratefully acknowledge Michael Acton from GL Noble Denton and Barbara Lowesmith from Hazard Analysis Ltd. for their acquisition and analysis of the original datasets along with productive discussions regarding follow-on interpretation of experimental results. REFERENCES [1] LaChance J, Tchouvelev A, Engebo A, Development of uniform harm criteria for use in quantitative risk analysis of the hydrogen infrastructure. Int J Hydrogen Energy, 2011;36:59-66. [2] Groth K, Zhu DF, Mosleh A, Hybrid Methodology and Software Platform for Probabilistic Risk Assessment. P Rel Maint S, 2008:413-8. [3] Coelho PJ, Numerical simulation of radiative heat transfer from non-gray gases in three-dimensional enclosures. J Quant Spectrosc Ra, 2002;74:307-28. [4] Richardson ES, Yoo CS, Chen JH, Analysis of secondorder conditional moment closure applied to an autoignitive lifted hydrogen jet flame. P Combust Inst, 2009;32:1695-703. [5] Auzillon P, Fiorina B, Vicquelin R, Darabiha N, Gicquel O, Veynante D, Modeling chemical flame structure and combustion dynamics in LES. P Combust Inst, 2011;33:1331-8. [6] Chamberlain GA, Developments in Design Methods for Predicting Thermal-Radiation from Flares. Chem Eng Res Des, 1987;65:299-309. [7] Johnson AD, Brightwell HM, Carsley AJ, A Model for Predicting the Thermal-Radiation Hazards from LargeScale Horizontally Released Natural-Gas Jet Fires. Process Saf Environ, 1994;72:157-66. [8] Sivathanu YR, Gore JP, Total Radiative Heat-Loss in Jet Flames from Single-Point Radiative Flux Measurements. Combust Flame, 1993;94:265-70. [9] API 521, Guide for Pressure-Relieving and Depressuring Systems, 4th edition, American Petroleum Institute, 1997. [10] Cook DK, Fairweather M, Hammonds J, Hughes DJ, Size and Radiative Characteristics of Natural-Gas Flares .1. Field Scale Experiments. Chem Eng Res Des, 1987;65:310-7. [11] Lowesmith BJ, Hankinson G, Acton MR, Chamberlain G, An overview of the nature of hydrocarbon jet fire hazards in the oil and gas industry and a simplified approach to assessing the hazards. Process Saf Environ, 2007;85:207-20.

[12]

[13]

[14]

[15]

[16]

[17]

[18] [19]

[20]

[21]

[22]

[23]

[24]

[25]

[26]

[27] [28]

[29]

Hankinson G, Lowesmith BJ, A consideration of methods of determining the radiative characteristics of jet fires. Combust Flame, 2012;159:1165-77. Schefer RW, Houf WG, Williams TC, Bourne B, Colton J, Characterization of high-pressure, underexpanded hydrogen-jet flames. Int J Hydrogen Energy, 2007;32:2081-93. Turns SR, Myhr FH, Oxides of Nitrogen Emissions from Turbulent Jet Flames - .1. Fuel Effects and Flame Radiation. Combust Flame, 1991;87:319-35. Molina A, Schefer RW, Houf WG, Radiative fraction and optical thickness in large-scale hydrogen-jet fires. P Combust Inst, 2007;31:2565-72. ISO 5167, Measurement of fluid flow by means of pressure differential devices inserted in circular crosssection conduits running full, International Standards Organization, Geneva, Switzerland, 2003. Ruggles AJ, Ekoto IW, Ignitability and Mixing of Underexpanded Hydrogen Jets. Int J Hydrogen Energy, 2012;(In Press). Measurement of Fluid Flow in Pipes Using Orifice Nozzle and Venturi. ASME MFC-3M-2004 ed. Wolfe WL, Introduction to Radiometry. Bellingham, WA: SPIE-International Society for Optical Engineering, 1998. Wayne FD, An Economical Formula for Calculating Atmospheric Infrared Transmissivities. J Loss Prevent Proc, 1991;4:86-92. Kalghatgi GT, Lift-off Heights and Visible Lengths of Vertical Turbulent Jet Diffusion Flames in Still Air. Combust Sci Technol, 1984;41:17-29. Birch AD, Brown DR, Dodson MG, Swaffield F, The Structure and Concentration Decay of High-Pressure Jets of Natural-Gas. Combust Sci Technol, 1984;36:249-61. Birch AD, Hughes DJ, Swaffield F, Velocity Decay of High-Pressure Jets. Combust Sci Technol, 1987;52:161-71. Ewan BCR, Moodie K, Structure and VelocityMeasurements in Underexpanded Jets. Combust Sci Technol, 1986;45:275-88. Yuceil KB, Otugen MV, Scaling parameters for underexpanded supersonic jets. Phys Fluids, 2002;14:4206-15. Harstad K, Bellan J, Global analysis and parametric dependencies for potential unintended hydrogen-fuel releases. Combust Flame, 2006;144:89-102. Molkov V, Special issue on hydrogen safety. J Loss Prevent Proc, 2008;21:129-30. Schefer RW, Kulatilaka WD, Patterson BD, Settersten TB, Visible emission of hydrogen flames. Combust Flame, 2009;156:1234-41. Karlsson B, Ribbing CG, Optical-Constants and Spectral Selectivity of Stainless-Steel and Its Oxides. J Appl Phys, 1982;53:6340-6.

11

Copyright 2012 by ASME

[30] [31]

Yolken HT, Kruger J, Optical Constants of Iron in Visible Region. J Opt Soc Am, 1965;55:842-&. McIlvaine JER, Barkaszi SF, Beal DJ, Anello MT, Laboratory Testing of the Reflectance Properties of Roofing Materials. Cocoa, FL: FSEC-CR-670-00, Florida Solar Energy Center, 2000.

12

Copyright 2012 by ASME

Das könnte Ihnen auch gefallen