Bullet Drag Aerodynamics

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/279961452

Bullet Drag Estimation with Compressible Conical Flow Equations and


Commercial Software

Research · July 2015


DOI: 10.13140/RG.2.1.4438.8965

CITATIONS READS
0 5,875

1 author:

Racha Lwali
California State University, Fullerton
3 PUBLICATIONS   0 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Racha Lwali on 10 July 2015.

The user has requested enhancement of the downloaded file.


Bullet Drag Estimation with Compressible Conical Flow
Equations and Commercial Software

Eric Larson1a, Racha Lwali1a, and Dejan Samardzic1a


a
Department of Aerospace Engineering and Engineering Mechanics, San Diego State University, San Diego,
California 92182-1308

Compressible conical flow theory is applied to the Spitzer-type tangent


ogive bullets along with two dimensional Prandtl-Meyer theory as it relates
to expansion waves. The theory is then compared to experimental data
obtained in a supersonic wind tunnel and to Solidworks Flow Simulator
commercial software results. A test section for the wind tunnel is designed
for drag data collection and the average theory deviation from collected data
is within 4.39%. Schlieren system is set up for shock wave angle
measurements. The averaged measured shock angles are within 2.39% of
the theory predicted values. Also, when compared to commercial software,
the theory approximates the drag coefficient to within 9.30% error while the
experimental data is within 5.42% of the software prediction. Although the
geometry warrants a detached shock due to the blunt tip, drag data is also
collected for a scaled model replica of the Sierra 0.308 bullet and compared
to both the Solidworks prediction as well as published results. In this case,
experimental data shows much higher deviation from software prediction
and literature. Mainly, the experimental data produced a 12.5% difference
from published results while Solidworks results remained within 3.35% from
the same published data.

Nomenclature
A = area θ = shock angle
γ = heat capacity ratio LC = cylinder length
= drag coefficient LF = frustum length
δ = Prandtl-Meyer geometry deflection LO = ogive length
D = drag force M = Mach number
d = maximum model diameter p = pressure
= base diameter σ = semi vertex angle
= rod diameter V = velocity

I. Introduction

B ullet aerodynamic properties play an important role in characterizing and predicting projectile
performance and further help in optimizing the application of rounds for the intended field of use.
Only in the recent years has there been a renewed interest in small caliber round aerodynamic testing [1].
This project aims at establishing an approximation method for obtaining the drag coefficient for Spitzer

1
Undergraduate Student, Department of Aerospace Engineering and Engineering Mechanics, San Diego State
University, 5500 Campanile Drive, San Diego, CA, 92182-1308

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
type bullets using the compressible conical flow theory combined with the two dimensional Prandtl-
Meyer theory where expansion fans are expected to be observed. These methods are explained in detail
in compressible flow literature most notably by Anderson [2] and Shapiro [3]. Spitzer-type models are
chosen due to the well-defined bullet geometry as well as their increased use in numerous applications
[4]. Figure 1 defines the geometry of a typical tangent type Spitzer bullet used in this experiment. The
center of the ogive radius is tangent to the intersection between the ogive and the cylinder.

Fig. 1 Geometry of a Spitzer-type tangent ogive bullet.

The five models that are tested differ in the length of the ogive radius. That is, the value 6D in Figure 1 is
varied from 5D to 9D for models A5 through A9 respectively. Notation A is assigned to the models to
differentiate them from model B7 which is a scaled replica of the Sierra 0.308 bullet used to compare the
experimental data with. The dimensions for the Sierra bullet are shown in Figure 2 bellow.

Fig. 2 Geometry of the Sierra .308 cal 155 grain PALMA MatchKing bullet (dimensions in
inches).

The B-type bullet indicates that the tip is flat as opposed to the A-type bullets. This definition is used for
clarity of data presentation in this paper and is not considered an industry standard. The models are
scaled to 3.26 times the diameter of the Sierra .308 round. This is done to achieve a maximum diameter
of 1”. The geometry definition for all the models can be found in the Appendix.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
The well-defined geometry of the bullet models allows for an analytic approach to drag calculations at
a high Mach number. The experimental data along with computational fluid dynamics analysis using
Solidworks Flow Simulator provide additional reference results for comparison with theory. Combined,
this information can prove to be a useful tool for projectile designers and a good starting point for future
experiments involving approximation methods of bullet drag.

II. Experimental Setup

A. Wind Tunnel
The testing is performed at the San Diego State University supersonic wind tunnel located in the
engineering building room E122. A Mach number of = 2 is used for all the tests and the stagnation
pressure at the inlet of the test section is at an average of = 30 . The tank pressure vessel is
brought to a pressure between = 130 to = 150 in order to obtain the desired Mach number
for approximately 10 to 15 seconds. The tunnel slides open at the location where the test section is
installed.

B. Test Section
The test section has a cross sectional area of = 36 without the data collection plate. Figure 3
shows the cross-section of the test section assembly with the data collection plate installed. The model is
screwed onto the rod which transfers the drag force onto the SB0-100 Transducer Techniques load cell.
The cell is secured to the 0.75 thick holding plate and is shielded on both sides by thin aluminum cover
plates (not shown). The entire assembly slides into the top and bottom plates of the rectangular test
section and is attached to the supersonic wind tunnel.

Fig. 3 Test section and data collection layout inside the tunnel.

The load cell has a maximum capacity reading of 100 and output accuracy of 0.01 . The
temperature effect on the load cell output is 0.08% of load/°F. A picture of the assembly outside of the
wind tunnel and without the bullet model attached can be found in Figure 4. The model can be seen
through two windows that are part of the tunnel layout. The windows allow for light to go through past
the model which is then projected for production of Schlieren images.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
Fig. 4 Test section prior to wind tunnel installation.

C. Models
The model geometry has already been noted in the introduction. The models are fabricated out of
aluminum and holes are threaded in the back to allow for model installation onto the holding rod. All of
the type A models have constant total and frustum lengths (see Fig. 1) of = 4.08 , = 0.59 .
Model B is trimmed to = 3.69 but its frustum length and overall diameter ( = 1.0 ) is equal to
the rest of the models. Figure 5 displays the A-type models.

Fig. 5 A-type bullet models with varying ogive radii.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
D. Test Procedure
Each model is installed onto the holding rod and the assembly is attached to the test section area. The
tunnel is then closed off and secured with holding pins. The pressure in the tank is allowed to build up
and, as the desired pressure is reached, the air is released through the tunnel. During the testing period,
the largest value of the drag is read off the Transducer Techniques board which is connected to the load
cell. A video of the Schlieren projection is filmed and the maximum stagnation pressure entering the test
section is recorded. The procedure is repeated for all the models and at least two readings are taken per
bullet.

III. Theoretical Analysis


The goal of this analysis is to obtain the drag coefficient of the A-type models. The experimental drag
of the B-type model is compared to the published literature. The effect of spin on rounds is neglected
since its contribution to drag is minimal compared to stability consideration [1]. Friction is also not
considered since the flow is at a high Reynolds number and the contribution to drag due to friction is
much less than the contribution due to wave drag [3]. The analysis is done for zero angle of attack and the
effect of shock interaction with test section walls, although a very important factor, is not considered in
this case. The drag coefficient is defined as [5]:

= (1)

The area in Eq. 1 is referenced to the maximum frontal area of the models, = , the Mach number is
= 2.0, = 1.4 and assumed constant, and the pressure is obtained from isentropic relations since the
Mach number and stagnation pressure entering the test section are both known. Therefore, the drag force
D is the unknown. Figure 6 defines the subscripts used for the step-by-step analysis.

Fig. 6 Subscripts denoting the areas of interest for theoretical analysis.

E. Total Drag
The total drag force can be found by integrating the static pressures in region 2, 4, and 5 over the
respective areas for which the normal vector is parallel to the flow:

= − ( − )− ( − ) (2)

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
where , , and are diameters of the overall model, model base, and the rod respectively. The
unknowns in Equation 2 are the static pressures acting on the corresponding areas.

F. Flow Over a Cone


Definition of conical flow is where all flow properties are constant along rays from the common
vertex. The equation describing the solution of conical flow is obtained with Taylor-Maccoll derivation
and given as [2]:

− − 2 + ( )+ − + =0 (3)

Equation 3 is an ordinary differential equation with as the only dependent variable. can be found to
with Equation 4:

= (4)

Equation 3 is solved numerically since there is no closed form solution. The shock angle is assumed with
a known Mach number. The unknowns are solved for and the iteration is continued until the final value
of θ is matched to the original guess and the velocity flow field is obtained.
NASA website [6] provides an interactive program which obtains the solution to Equation 3 using the
same Taylor-Maccoll analysis. From this, the stagnation pressure drop, shock angle, Mach number along
the surface of the cone, and the pressure ratio on the surface with respect to the free stream pressure are
obtained and used in this paper.

G. Shock Angle
The shock angle is predicted using the above mentioned analysis with a known Mach number and
model geometry. It is assumed that the flow reacts to the semi vertex angle of the bullet in the immediate
vicinity of the tip. This angle is, by the definition of geometry, equivalent to the angle that the ogive
radius sweeps from the cylinder body to the tip of the bullet and it is presented in the Appendix along
with the model geometries. Let the shock wave angle be denoted by in accordance with Equation 3 and
the model semi vertex angle be denoted by . The experimental shock wave angle is measured by taking
a still-shot image of Schlieren videos and then imported into AutoCad software. The shock wave front
and the bullet centerline are outlined and the angles measured within the software itself.

Fig. 7 Method of measuring experimental shock wave angles with use of AutoCad software.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
The image is, of course, imported to a 1: 1 scale. To illustrate this process, the imported image for model
A9 is shown in Figure 7 along with the corresponding angles. The values are then averaged and
compared to theory.

H. Ogive Static Pressure


Since the analysis of the models involves approximating the ogives as conical shapes, Taylor-Maccoll
theory predicts a constant Mach number along the surface of the cones as well as a constant static
pressure. But the ogive semi vertex angle as referenced to the incoming flow varies from its maxium at
the tip to a value of zero at the ogive connection to the cylinder body. Therefore, in order to avoid an
overestimate for the surface pressure and the surface Mach number by using the semi vertex angle in the
vicinity of the bullet tip, an average angle ′ is used as the input variable for the analysis. This is the
angle that the ogive would have if it were converted into a perfect cone by connecting the vertex directly
to the cylinder. In other words;

= (5)

where is the ogive length and is the bullet diameter. Using ′ and M as the input variables, the static
pressure acting on the ogive surface is then easily obtained from the NASA program outputs as
previously discussed.

I. Expansion Fans
Remaining with the assumption of conical geometry approximation, the flow now passes through
expansion fans as it is redirected by the same angle ′ to its original free stream direction. The stagnation
pressure remains equal, that is, = throughout the process [7]. The flow in this region may be
considered plane. The changes in radius in this location are negligible when compared to the overall
radius itself and so the flow in the immediate neighborhood is given exactly by the two-dimensional
relations for the Prandtl-Mayer flow [3]. For a calorically perfect gas, Equation 6 can be used to obtain
the Mach number downstream of the expansion fans.

= ( )− ( ) (6)

Tabulated values of ( ) can be found in Anderson’s [7] Appendix C. The value of is equal to ′ for
expansion fans between regions 2 and 3. With the known stagnation pressure and Mach number for the
region, isentropic relations can be used to determine the static pressure on the surface.
Similar reasoning is used in transferring the flow from region 3 to 4 with the exception that = 9° for
all models. Region 3 is of no interest to the drag contribution since, as previously stated, the effects of
friction are neglected in this paper.

J. Base Pressure
The pressure in region 5 is determined through experimental data since there is no complete theory for
the mechanics of the wake formed at the base [8]. Using the summary of measurements of base pressure
versus Mach number (after Kurzweg) published in Ref. [3] an approximate value can be found for
= 0.5. Note that in this case, the base pressure is referenced to the free flow static pressure. All the
unknowns in Equation 2 have now been defined.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
IV. Results

K. Shock Angles
Shock angles obtained through experiments and theory are plotted in Graph 1 as functions of the ratio
between the total model length and the ogive radius. As expected, the higher ogive radius bullets are
accompanied with a lower shock angle. The two sets of data show a higher divergence in results for
models with a lower ogive radius but both sets display a nearly linear trend.

Oblique Shock Angles vs. Non-Dimensional Model Length


45.0

44.0

43.0
Shock Angle, θ [deg]

42.0

Experiment
41.0
Theory

40.0

39.0

38.0

37.0
0.400 0.500 0.600 0.700 0.800 0.900
Lt/R

Graph 1 Experimental and theoretical shock angles plotted versus non-dimensional model length.

The experimental results are on average 2.39% lower than what the theory predicts with the highest
error of 3.68% with model A6 and the lowest error of 0.37% with model A9. Measured and predicted
results are summarized in Table 1.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
Model θave θtheory
- deg deg
A5 42.6 43.9
A6 40.1 41.6
A7 38.6 39.8
A8 37.7 38.4
A9 37.4 37.5

Table 1 Summary of averaged measured and predicted shock angles.

L. Drag
The drag coefficient, as defined in Equation 1, is plotted in Graph 2 as a function of the non-
dimensional model length. The CFD data set represents the results obtained through SolidWorks Flow
Simulation analysis. The data points for model B7 are also plotted for values obtained through CFD,
experiment, and literature [4].

Drag Coefficient vs. Non-Dimensional Model Length


0.390

0.370

0.350 CFD
Drag Coefficient, cD

Theory

0.330 Experiment
B7 CFD

0.310 B7 Experiment
B7 Literature

0.290

0.270
0.390 0.490 0.590 0.690 0.790
Lt/R

Graph 2 Drag coefficient versus the non-dimensional model length.

It is seen that, as the ogive radius is decreased, the drag coefficient is increased. This is, once again, to
be expected due to the higher model tip angles for lower ogive radii. The experimental data follows the
CFD prediction for models with a higher ogive radius while it approaches the theoretical analysis for the
lower ogive radius models. Both theory and CFD results follow a similar trend line but they are offset by
an average bias in the drag coefficient by approximately ∆ = 0.04.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
The average drag coefficient theory error, when compared to the CFD analysis, is 9.30% and 4.39%
when compared to experimental results. The collected data displays an error of 5.42% from CFD
analysis. It is immediately apparent from data in both Graphs 1 and 2 that the higher radius ogive models
agree much better in theory, experiment, and CFD analysis than the models with lower radius ogive. This
is expected because a model with a higher ogive radius closer represents a conical shape than the one with
a lower ogive radius. Most of the measured data falls in between the theory and the CFD analysis with
the exception of the A5 model which recorded a 1.07% higher reading when compared to theory.
The drag for model B7 was collected from experimental data and CFD analysis and then compared to
the literature. The theoretical analysis is not performed since conical flow equations cannot be reasoned
to provide an adequate model with a detached bow shock. It can be seen from Graph 2 that the data for
model B7 does not provide a clear result for drag prediction. The experiment data produced a drag
coefficient higher than published literature by 14.3% and also higher than CFD analysis by 18.3% while
the literature and CFD data are within a 3.35% difference. In addition to the total drag analysis for the
B7 model in CFD, a pressure distribution plot is obtained and presented in Figure 8.

Fig. 8 CFD pressure plot distribution for bullet model B7.

The highest value in the accompanying legend is 17.84 (red) and the lowest is 1.75 (blue)
with average difference between the increments on the scale of 1.24 . The high pressure region is
located in the vicinity of the flat tip due to the normal shock. The low pressure region is apparent in the
bullet base as expected. Figure 9 is a Schlieren image of the model B7 inside the supersonic wind tunnel
and it is presented here to show the similarities between the experimental and CFD data. The difference
between models A7 and B7 is predicted to be 6.36% by CFD while experimental data shows an error of
24.9% between the models. It should be noted that in both experimental and CFD analysis, the drag on
the B7 model is lower than the drag on model A5.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
Fig. 9 Schlieren image of the bullet model B7.

M. Error Sources
All of the experimentally obtained shock angles are lower than the theoretically produced data. In
theory, it was assumed that the angle that the flow sees is equal to the angle of the bullet in the immediate
vicinity of the tip. But the angle of the geometry varies along the length of the ogive and so the flow
interacts with a larger portion of the geometry than just the tip of the model. This is why, in calculation
of the drag coefficient, the average angle ′ is used rather than the tip angle .
The CFD model analysis produced a much lower estimate of the drag coefficient than the theory and
experimental data. The difference in the CFD modeling is that the bullets in SolidWorks do not include
the holding rod and the holding plate. Since the rod does not see any normal flow, this pressure
contribution is not accounted for in the CFD analysis but it is subtracted in theoretical setup by the
differences in areas in Equation 2. The model Schlieren images in Figure 7 and in the Appendix show
numerous shock interactions from the test cell walls and the holding plate. None of these factors are
accounted for in theoretical or CFD analysis.
The temperature in the pressure tank increased with increased testing and the high temperature
differential produced outlying data points for identical models in certain cases. The time interval between
the tests was then increased in order to allow the tank to return to ambient temperature and allow for more
consistent data collection.
Moisture was present in the test section for various runs and it produced the largest deviations in the
readings. Figure 10 shows a Schlieren image of model A8 immediately after the test run. This run
produced an error deviation from the previous averaged runs for this same model that is higher by 35.7%.
The value for this run was reasonably considered an outlier.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
Fig. 10 Model A8 Schlieren image with test cell moisture buildup.

V. Conclusion
This project produced bullet drag data with some well-defined trends and it supported some of the
initial expectation of final results. Mainly, an increase in drag was observed for models with decreased
bullet ogives. The conical flow theory can be said to apply as an appropriate initial estimate of the upper
bound of the drag coefficient for models in which the radius ogive is less than six times the length of the
bullet diameter. Similarly, the CFD analysis can be used as the lower bound for the drag coefficient
estimates.
In future testing, a much smaller holding plate could be designed by the use of miniature button-type
load cells to minimize the shock interaction with the holding assembly. Total and static temperature
probes, installed immediately before the test cell entrance, could increase accuracy of the data analysis by
allowing for temperature corrections. Also, a static pressure probe could be installed in the same location
in order to verify the Mach number entering the test cell since the Mach number is assumed constant due
to the wind tunnel geometry. Moisture control in the tunnel could be improved or instrumentation for
measuring the amount of moisture could be installed. Bullet models that would produce a much higher
drag differential readings should be designed in order to compensate for the limited precision of current
instrumentation.
A more extensive experimentation should be performed on B-type models to compare them with their
A-type counterparts and also to determine the drag differential between the B-type models as compared to
the differential of the A-types. A test cell capable of providing spin to the models would allow for
determination of the drag influence due to the various projectile spin velocities. And finally, a larger
database of Schlieren images could support a research project to determine the equivalent semi-vertex
angle of an ogive that produces a given shock angle.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
Appendix

Model D d LF LC LO LT R σ
- in in in in in in in Deg. (°)
A5 1.00 0.816 0.588 1.3097 2.1794 4.0771 5.00 25.84
A6 1.00 0.816 0.588 1.0912 2.3979 4.0771 6.00 23.56
A7 1.00 0.816 0.588 0.891 2.5981 4.0771 7.00 21.79
A8 1.00 0.816 0.588 0.7052 2.7839 4.0771 8.00 20.36
A9 1.00 0.816 0.588 0.5311 2.958 4.0771 9.00 19.19
B7 1.00 0.816 0.588 0.891 2.2137 3.6927 7.00 18.44

Table A1 Model geometries.

Model M1 P01 P1 σ1 θ1 θ2 θave θtheory σ' P02/P01 P02


- - psi psi deg deg deg deg deg deg -
A5 2 30 3.834 25.8 42.2 42.9 42.6 43.9 12.9 0.999 29.97
A6 2 30 3.834 23.6 39.6 40.5 40.1 41.6 11.8 0.999 29.97
A7 2 30 3.834 21.8 37.8 39.3 38.6 39.8 10.9 0.999 29.97
A8 2 30 3.834 20.4 37.5 37.9 37.7 38.4 10.2 0.999 29.97
A9 2 30 3.834 19.2 36.2 38.5 37.4 37.5 9.6 0.999 29.97

Table A2 Data from theoretical analysis.

Model P2/P1 P2 ν2 ν3 P3 δ2 ν4 ν5 P4/P3 P4 P5/P1


- - psi deg deg psi deg deg deg - psi -
A5 1.469 5.632 18.7 31.6 2.804 9.00 31.73 40.73 0.552723 1.549834 0.55
A6 1.395 5.348 20.2 31.9 2.593 9.00 33.02 42.02 0.557512 1.445629 0.55
A7 1.356 5.199 20.7 31.6 2.804 9.00 31.73 40.73 0.552723 1.549834 0.55
A8 1.310 5.023 21.3 31.5 3.031 9.00 30.43 39.43 0.578727 1.754122 0.55
A9 1.284 4.923 21.9 31.5 3.031 9.00 30.43 39.43 0.578727 1.754122 0.55

Table A3 Data from theoretical analysis (cont.).

Drag Data [lbs]


A5 3.3 3.05 -
A6 2.97 2.81 -
A7 2.57 2.35 2.67
A8 2.26 2.5 2.7
A9 2.42 2.44 -
B7 3.2 3.11 -

Table A4 Raw experimental data.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
Fig. A1 Model A5 Schlieren image.

Fig. A2 Model A6 Schlieren image.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
Fig. A3 Model A7 Schlieren image.

Fig. A4 Model A8 Schlieren image.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
Fig. A5 Model A9 Schlieren image.

Fig. A6 Test cell data (see Model SB0-100)

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
Acknowledgments
The authors are grateful to the Mechanical and Aerospace Engineering department support staff and to
the East County Metal Supply company for their generous donation of materials for testing equipment
fabrication.

References
[1]Sidra, I. S., and Bradley, E. H “Aerodynamics and Flight Dynamics Characteristics of 5.56-mm Ammunition:

M855,” Army Research Laboratory, ARL-TR_5182, May 2010.

doi: 10.2514/3.13046

[2] Anderson, J. D. Jr, Modern Compressible Flow with Historical Perspective, 2nd ed., McGraw-Hill, New York,

1990, Chap. 10.

[3] Shapiro, A. H., Terster, W., The Dynamics and Thermodynamics of Compressible Fluid Flow, Vol. 2, The

Ronald Press Company, New York, 1954, Chap. 17.

[4] Litz, B., Applied Ballistics for Long-Range Shooting, 2nd ed., Applied Ballistics LLC, 2011.

[5] Cengel, Y. A. and Cimbala, J. M., Fluid Mechanics: Fundamentals and Applications, 2nd ed., McGraw-Hill,

New York, 2010, Chap. 7.

[6] ShockSim Version 1.3e, National Aeronautics and Space Administration, Retrieved Dec 1, 2012, from

http://www.grc.nasa.gov/WWW/k-12/airplane/shock.html

[7] Anderson, J. D. Jr., Fundamentals of Aerodynamics, 4th ed., McGraw-Hill, New York, 2007, Chap. 9.

[8] Liepmann, H. W. and Rosko, A., Elements of Gasdynamics, Dover Publications, New York, 2001, Chap. 4.

PDF compression, OCR, web optimization using a watermarked evaluation copy of CVISION PDFCompressor
View publication stats

You might also like