You are on page 1of 8

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/286244002

Experimental Verification of Drag Forces on


Spherical Objects Entering Water

Article January 2014


DOI: 10.4172/2324-8661.1000126

CITATIONS READS

2 27

2 authors, including:

Thomas Shepard
University of St. Thomas
29 PUBLICATIONS 29 CITATIONS

SEE PROFILE

All content following this page was uploaded by Thomas Shepard on 23 February 2016.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
Abraham J Mar Biol Oceanogr 2014, 3:2

Journal of Marine
http://dx.doi.org/10.4172/2324-8661.1000126

Biology & Oceanography


Research Article a SciTechnol journal

they fall through the water and consequently, the measured ocean
Experimental Verification temperatures. For the application of climate monitoring, ocean heat
content is by far the largest thermal reservoir and uncertainty of
of Drag Forces on Spherical ocean measurements are the largest climate monitoring uncertainty.

Objects Entering Water Since the second half of the 20th century, millions of XBT devices
have been dropped into ocean waters from various ocean-going
John Gorman1, John Abraham2*, Dillon Schwalbach2, Thomas vessels. In some cases, the drops are part of carefully controlled
Shepard2, John Stark2 and Franco Reseghetti3
experiments where the XBT probes are released from specified heights
above the ocean (2-3 meters). On the other hand, in other cases, the
XBTs are dropped from commercial shipping vessels from heights
Abstract
Objects which pass from gas regions to liquid regions experience
that may be ~30 meters. As a consequence, the impact velocity of the
elevated impact forces associated with the acceleration of device at the water surface differs widely depending on the specific
the surrounding liquid. In order to investigate these forces, nature of the experiment.
complementary experiments and simulations were performed on a
For the XBT device, depth is not measured directly, rather it
sphere that traveled from air to water with an impact velocity of 2
m/s. It was found that the two methods gave results that were in is inferred from experimentally obtained correlations of time and
very good agreement. In particular, the depth vs. time trajectory depth. The correlating equations generally take the form shown in
of the sphere closely matched. A fitted polynomial allowed the Equation (1).
entry region acceleration to be extracted. The two independent
accelerations were found to be in excellent agreement. A simple Depth = at2 + bt + c (1)
momentum analysis allowed extraction of drag coefficients and they The coefficients in Equation (1) characterize the rate of descent.
too were in very close agreement. The values of the drag coefficient
They are specific to the type of probe, including its size and weight. A
also agree with prior studies. Finally, a comparison of the width
of the splash region provided further validation of the methods.
detailed discussion of fall rate equations is provided in the literature
Taken together, the excellent agreement lends strong support to and information about the coefficients is given in [1-19]. A historical
the numerical methods which were employed. Entry-region drag review of oceanographic measurement systems with particular
coefficients enable users to predict the trajectory of devices in other emphasis on the XBT device is provided in the literature [13]. In
applications. Those applications are not limited to the object size the literature cited here, discussions on both depth bias and fall rate
and impact velocity that were studied here. equations can be found.
Keywords There are some obvious issues with the fall rate equation because
Climate change; Ocean heating; Drag forces; Entry forces; Water it cannot be used to account for the specific details of the launch. For
entry; Ocean temperatures; Climate monitoring instance, launch height and water temperature, which are known to
affect fall rate, are not accounted for explicitly. As a result, fall rate
Introduction equations obtained from carefully controlled drops of ~2 m in warm
tropical waters are used for much higher drop heights in all water
Many engineering, scientific, military, and commercial enterprises temperatures. The differences in operating conditions can lead to
involve the transfer of solid objects from gaseous to liquid regions. depth errors that are a significant uncertainty in closing the Earths
Some examples include the release of oceanographic measurement energy balance [13].
devices into water, missiles or torpedoes which enter the water, the
impact of boat hulls on water surfaces, and others. Another drawback with the form of Equation (1) is that is
preordains a constant acceleration which is not a priori expected,
During the penetration of the liquid surface, large impact forces particularly in the entry region.
are realized on the solid object. The genesis of the forces is associated
with the transfer of momentum from the solid object to the liquid and While the ultimate motivation for this study is the development
the subsequent acceleration of liquid in front of the object. There may of a numerical methodology which can be employed for XBT devices,
also be extra drag forces associated with the formation of a gas cavity the shape studied here is more simple. The sphere was chosen for the
behind the projectile. present geometry because it has the most extensive literature available.
The studies date back more than 100 years and only an abbreviated
It is important to quantify these impact forces because they greatly history is discussed here. Interested readers are referred to [19] which
affect the trajectory of the solid object after impact. For instance, for more fully explores the rich literature on the sphere-entry problem.
oceanographic measurement devices, like the eXpendable Bathy
Thermograph (XBT), the impact forces can affect the rate at which An early paper that explored impact forces was written by
Richardson EG (1948) [20] which determined pressure variations
on surfaces as they impacted a water body. Three contemporary
*Corresponding author: John Abraham, School of Engineering, University of
studies [21-23] developed analytical methods to predict the forces on
St. Thomas, 2115 summit Ave, St. Paul, MN 55105-1079, USA, Tel: 651-962-
5766; Fax: 651-962-6419; E-mail: jpabraham@stthomas.edu objects during the early stages of impact by balancing the momentum
transferred between the moving object and the stationary fluid.
Received: November 11, 2013 Accepted: January 24, 2014 Published:
February 05, 2014

All articles published in Journal of Marine Biology & Oceanography are the property of SciTechnol, and is protected by
International
InternationalPublisher
PublisherofofScience,
Science, copyright laws. Copyright 2014, SciTechnol, All Rights Reserved.
Technology
Technologyand
andMedicine
Medicine
Citation: Gorman J, Abraham J, Schwalbach D, Shepard T, Stark J, et al.,(2014) Experimental Verification of Drag Forces on Spherical Objects Entering Water.
J Mar Biol Oceanogr 3:2.

doi:http://dx.doi.org/10.4172/2324-8661.1000126

The following decades produced a series of government sponsored literature information.


research with applications to naval ordinance trajectories. Those
Despite this level of agreement, it was found that the experimental
studies encompassed both experimental and analytical methodologies
data was not existent for depths beyond approximately one radius.
and expanded beyond spheres to other relevant shapes such as ogives,
Traditionally, depth is presented in terms of a dimensionless distance
cylinders, and cones [24-28].
b (Equation (4)) where
More recently, new studies looking at high-speed entry problems penetration depth
b= (4)
or new analysis methods have been published with [29,30] as sphere radius
examples. Others evaluated the impact of spinning motions of the The fully submerged condition corresponds to b > 2.
sphere [31] and issues related to optical measurement accuracy [32].
To overcome the paucity of literature information beyond
The reference list here is not intended to be exhaustive, readers are
those of [20-30], new experiments have been performed. With the
invited to go through reference [19] for a deeper discussion.
assistance of high-speed photography, the fall rate of a sphere will
The Computational Method be determined in the entry region (until b ~ 2). In addition, the
calculations of [19] have now been extended to the same depth.
Dynamic model of device trajectory Comparison between the simulations and the experiments will be
In order to overcome the deficiencies in the standard fall rate shown to be mutually reinforcing. The comparisons will be made
equation, dynamic models have been developed [9-12,17]. Dynamic for both the depth of the sphere and the shape of the water splash.
models are not limited to XBT devices, rather they can be applied to Details of the numerical simulation are reported in [19] and will not
any objects experiencing drag, buoyancy, and gravity forces. These be repeated here for brevity. The simulations and experiments are
models apply a force-momentum balance to objects falling through matched in all respects, including the size and mass of the sphere, the
the water column which is expressed mathematically as entry velocity, and the fluid types and temperatures. The consistency
between the two independent investigations allows clear conclusions
d (mV )
= Fdrag + Fbuoyancy + Fweight (2) to be drawn regarding the accuracy of both methods.
dt
The left side of Equation (2) is the rate of change of linear
Details of present simulation domain
momentum of a device. On the right are drag, buoyancy, and weight
forces, respectively. In order to apply this model to calculate the While the details of the simulation are discussed fully in reference
trajectory of a device underwater, it is necessary to know the drag [19], here it will be noted that the mass of the sphere was 0.0335 kg
forces. Drag forces exist between any solid object that has a relative and the radius was 0.02 m. The density of the sphere was identical to
velocity with respect to the fluid surrounding it. Drag is caused by that of water. For both the simulations and experiments, the sphere
both friction between the object and the fluid and by pressure forces impacts the water surface at 2 m/s.
(pressures tend to be higher in the front of an object and lower to its
The solution domain is shown in Figure 1. It is not drawn to scale
rear).
but nevertheless shows that the external boundary conditions were
Most commonly, the drag force is found by multiplying a drag set far enough from the moving sphere so that they did not influence
coefficient (Cd) by the dynamic pressure of the fluid and the frontal the desired results.
area of the object, so that
The solution made use of the Shear Stress Transfer model of [33].
Fdrag = Cd 12 AV 2 (3) This model accounts for turbulence through an eddy viscosity which
is incorporated into the Reynolds Averaged Navier Stokes equations.
Here, , A, and V are the fluid density, the projected area of the
Details of the solution method and the computational mesh are
object (R2) , and the object velocity, respectively.
provided in references [19,33].
It still remains to find values of the drag coefficient that
allow utilization of Equation (3). This term can be determined by Details of the experiment
experiment or by numerical simulation using the methodology Experiments were conducted using the equipment and setup
termed computational fluid dynamics (CFD). This latter approach illustrated in Figure 2. A high-speed camera (Photron FASTCAM
was taken in a series of studies on XBT underwater trajectory [9- SA1.1) outfitted with a Nikon 50 mm lens was positioned so that the
12] and the results matched experiments. In other studies [17], drag center pixel of the sensor aligned with the intersection of the water
coefficients were estimated based on similar shaped objects. surface. A 279.4 381 762 mm 75.7 L plastic tank was modified to
A major shortcoming in all of these works, including [9-12,17] include two opposing polycarbonate windows allowing unobstructed
is that the drag coefficients are based on submerged devices. While view through the tank. For the experiments, it was filled with 57 L of
this assumption holds true for the vast majority of the fall, it is not tap water. Three lights (Lowell DP #ML 86) were diffused through a
appropriate at the water surface where very high impact forces may sheet of tracing paper and positioned to backlight the falling sphere.
exist. Without accounting for impact forces, it has been shown that These conditions allowed the camera to capture the falling sphere
drop height has a large impact on the ultimate fall rate of the probe as a black silhouette (Figure 3), minimizing aberrations caused
[12]. by shadows or reflections. Both the top and bottom surfaces were
captured simultaneously.
A recognition of this shortcoming has motivated new numerical
simulation studies such as [18,19]. In [19], a numerical simulation The sphere was released from a specific drop height by means of a
was performed for the motion of a spherical object that passed from funnel and vacuum suction positioned above the middle of the tank. A
an air space to a water region. That data allowed verification of the vacuum pump (JB #DB-85N) pulled a constant suction on the sphere
simulations of [19] which were found to very closely match the until the line was released to atmospheric pressure, allowing the

Volume 3 Issue 2 1000126 Page 2 of 4


Citation: Gorman J, Abraham J, Schwalbach D, Shepard T, Stark J, et al.,(2014) Experimental Verification of Drag Forces on Spherical Objects Entering Water.
J Mar Biol Oceanogr 3:2.

doi:http://dx.doi.org/10.4172/2324-8661.1000126

Figure 1: Solution domain including air and water regions and solid sphere moving vertically downwards, reproduced from [20].

sphere to drop instantaneously and without any externally induced Table 1: A summary of the measured sphere properties.
forces. The camera was set to capture at a rate of 5,400 frames per Mass Diameter
Ps
second (fps) and a resolution of 1024 1024 pixels using the software g g (95% CI) Mm mm (95% CI)
Photron FASTCAM Viewer, version 3.0. Pw
33.428 0.0005 39.815 0.1012 1.013
The sphere used in the experimental portion of the study was
made to match the weight and diameter of the sphere used in the
Paper A wt. sandpaper. After sanding, the ball was rinsed using tap
simulation. A celluloid sphere (ping pong ball, Prince #PB-TG6) had
water and dried. A final coat of WX2100 was applied to the entire
an 11.1 mm hole drilled into it which was then filled with molten
sphere. Careful surface preparation provided a very smooth surface
paraffin (McMaster-Carr #1085K91). The wax was allowed to
so that surface roughness was expected to be negligible. Similarly, the
harden and a 7.94 mm hole was drilled down to the center of the
simulations were carried out with a smooth-surface model. Once the
sphere. A 9.525 0.127 mm diameter 101-1020 carbon steel sphere
final coat had dried, the sphere was handled using nitrile gloves and
(McMaster-Carr #96455K54) was press-fit into the sphere center and
stored in a closed container. The sphere properties are summarized
the remaining void was filled with molten paraffin wax. The sphere
in Table 1.
surface was washed thoroughly and then coated with WX2100, a
hydrophobic surface coating. After drying for at least 24 hours, The mass of each sphere was measured using a Mettler-Toledo
flaws in the coating were removed using 3M 413Q 400 Wetordry MS303S digital scale. The diameter of each sphere was measured

Volume 3 Issue 2 1000126 Page 3 of 4


Citation: Gorman J, Abraham J, Schwalbach D, Shepard T, Stark J, et al.,(2014) Experimental Verification of Drag Forces on Spherical Objects Entering Water.
J Mar Biol Oceanogr 3:2.

doi:http://dx.doi.org/10.4172/2324-8661.1000126

Vacuum Line
Funnel

Sphere

Camera
Light Source Figure 5: Somparison of simulated and experimental sphere depths.

Water Tank
specific gravity were made using a Chase hydrometer. A Cannon
#50 capillary viscometer was used to measure the viscosity of the
water. Surface tension was measured using a CSC-DuNuoy 70535
Surface Tensiometer. Finally, A Fluke 52II thermometer and Type
Figure 2: Schematic of the experimental setup.
K thermocouples measured water temperature in the tank at a depth
of 40 mm. The results from these measurements are summarized in
Table 2.
During each experimental run the sphere was loaded into the
funnel with the filled hole oriented such that it would be on the
trailing end during impact. The funnel was dried between runs to
ensure that no residual water would be transferred to the ball prior
the run. The sphere was dropped from a height of 222 mm measured
from the bottom of the sphere to the surface of the water.
The sphere position vs. time data was obtained by an algorithm
developed in MATLAB which would read the series of images of the
sphere as it fell. The specific MATLAB commands used were im2bw
which converted a grayscale image into a binary image based on an
Figure 3: Sequential photographs of the sphere prior to and during impact. user-defined intensity threshold and regionprops (Extrema) which
identifies objects in the binary image and gathers position information
regarding the edges of the object. As the sphere approached the water
interface from above the bottom of the sphere became obscured by
the water interface for roughly 7 frames (Figure 4a). Thus the position
of the top of the sphere was used in determining its velocity at impact.
Upon impact, the bottom of the sphere was clearly seen emerging
below the air-water interface in Figures 4b and 4c. The position of
the bottom of the sphere upon impact was used in determining the
depth vs. time data in the experimental portion of this study as it was
not distorted by interference of splashing water. The bottom position
was shown to reliably track with the top position of the sphere until
Figure 4: Sequential photographs of sphere impacting water surface. the splash obscured the top position. It is noted that upon entering
the water the sphere appears magnified so a different pixel/mm
Table 2: A summary of the measured water properties. scaling ratio was used for the air and water. For data analysis the zero
Temperature Specific gravity Viscosity Surface Tension position and zero time correspond to the undisturbed free surface of
C C (95% CI) (95% CI) mm2/s2 mm2/s2 (95% CI) dyne/cm dyne/cm (95% CI) the water and the moment of impact.
22.6 0.050 0.998 0.001 1.121 0.0121 71.5 0.480
A total of six runs were conducted in order to determine
using a Mitutoyo caliper, model #CD-6CSX. A total of six diameter the experimental repeatability which played the largest factor in
measurements were taken at a position offset by 45 relative to all determining experimental uncertainties as the relative uncertainty
other measurements. in pixel location and time was negligible. The impact velocity was
measured as the average velocity within 10 mm of the air-water
Repeated measurements of the properties of the water were interface and determined to be 1.96 m/s 1%. The splash width
taken both before and after all experiments. Measurements on water was determined experimentally from the images for the six trials

Volume 3 Issue 2 1000126 Page 4 of 4


Citation: Gorman J, Abraham J, Schwalbach D, Shepard T, Stark J, et al.,(2014) Experimental Verification of Drag Forces on Spherical Objects Entering Water.
J Mar Biol Oceanogr 3:2.

doi:http://dx.doi.org/10.4172/2324-8661.1000126

datasets, it is possible to calculate an entry-average drag coefficient by


rearranging and averaging Equation (2) so that

m(a + g ) (6)
Cd = average
1
AV 2
2

throughout the entry where a represents the entry-averaged


acceleration which is defined as positive upwards.
Entry-averaged accelerations are determined by twice
Splash Width differentiating and averaging Equations (5a) and (5b) with respect to
time. That operation gives equations (7a) and (7b)

aexperiments = 13.79 m / s 2 (7a)


and

asimulations = 13.54 m / s 2 (7b)


The corresponding entry-averaged drag coefficients are found to
be
Cd = 0.42 from experiments (8a)
and
Cd = 0.39 from simulations (8b)
These results are within 5% of those of [20] and strongly confirm
Figure 6: Definition of the splash width used to compare simulated and the prior results.
experimental results.
An alternative definition of the drag coefficient is that which can
be obtained by using the initial velocity at impact and the average
corresponding to a sphere submersion depth of 40 mm (i.e b = 2). The acceleration during the entry process. That definition is
uncertainty in the experimental splash width was 1%. To estimate m( a + g )
the uncertainty in the position vs. time data, the position data for the C 'd = (9)
1
AVimpact
2
six runs was averaged at a given time step. A comparison of individual 2
runs to the averaged data showed that the average deviation between
a single data point and the averaged data points was 0.5%. Here, the velocity in the denominator is taken to be 2 m/s. With
this definition, it is found that
Results and Discussion Cd=0.314 from experiments (10a)
Entry region drag and
The result of the just described experiments is a depth-time Cd=0.311 from simulations (10b)
relationship which is shown alongside its numerical counterpart in
Figure 5. There, the two curves are seen to show increasing depth The utility of Cd is that in typical applications, only the impact
with time; both display a deceleration of the sphere. To quantify the velocity is known so it is difficult to employ Equation (6) directly. On
motion, a third-order polynomial has been fit through both sets of the other hand, Equation (9), with the appropriate value of the initial
results. The highest-order terms in the polynomial provide time- velocity and Cd , can be used to determine the entry-region-averaged
varying deceleration. As stated earlier, prior expressions using second acceleration.
order polynomials would enforce a constant acceleration within the It has already been noted here, in reference [19], and in other
entry region-there is no reason to expect such an occurrence. work, the drag coefficient was shown not to depend on the size of the
The resulting polynomials are: sphere or its impact velocity. Consequently, the results provided here
in Equations (8a), (8b), (10a) and (10b) can be used for other sized
Depthexperiments = 111.38 t 3 10.235 t 2 + 2.0074 t (5a) spheres and for other impact velocities.
from experiments, with the depth in meters and t in seconds.
Entry region splash patterns
Similarly
Another useful comparison relates to the splash pattern and the
Depthstimulations = 233.4 t 3 13.77 t 2 + 1.937 t (5b)
size of the trailing cavity. One metric of interest is the width of the
from simulations. cavity zone when the top of the sphere is coincident with the surface
of the undisturbed fluid. The cavity width is measured at the liquid
It can be seen that the experimental results fall slightly faster than surface. Figure 6 has been prepared which shows the simulated
their simulated counterparts. sphere passing into the water with a trailing air region. The air-fluid
From Figure 5, it is seen that the simulation and experiments regions are colored by different shades of grey. For the simulated case,
provide very good agreement (approximately 5%). From both the splash width was found to be 0.048 m (4.8 cm).

Volume 3 Issue 2 1000126 Page 5 of 4


Citation: Gorman J, Abraham J, Schwalbach D, Shepard T, Stark J, et al.,(2014) Experimental Verification of Drag Forces on Spherical Objects Entering Water.
J Mar Biol Oceanogr 3:2.

doi:http://dx.doi.org/10.4172/2324-8661.1000126

and the impact velocity. Also, the drag coefficient dependence on


dimensionless depth was found to be independent of initial velocity
over the range of penetration depths provided.

Conclusion
The present study focused on experimentally verifying a
numerical method that was set forth in a prior paper. Both the
numerical and experimental investigations were performed with
all relevant parameters equal (mass, impact velocity, fluids, sphere
size, etc.). The studies were carried out until the top of the sphere
was coincident with the surface of the undisturbed water. High-speed
camera images allowed virtual continuous tracking of the sphere at all
instances in time. This data was used to obtain instantaneous velocity
and acceleration. From a momentum balance, the drag forces and
drag coefficients were found.

Figure 7: Photograph of splash when the sphere top is coincident with the Similarly, from the simulations, the timewise variations of
undisturbed liquid surface. position, velocity, acceleration and force were found. A comparison
of the two independent results showed that the predictions of the
spheres were in very close agreement. It was also found that the entry-
averaged drag coefficients were in excellent agreement. They also
agreed with prior publications which were validated by experiments
carried out by other researchers.
A second comparison was made on the fluid dynamic flow
patterns. In particular, the splash width was defined as the width
of the cavity region at the top of the sphere when the sphere top is
coincident with the undisturbed water. It was found that the simulated
and experimental splash widths agreed to approximately 6 %.
From these studies, it is possible to determine the fall-rate
characteristics of spheres with other sizes and impact velocities
during their penetration into a liquid water surface.
References
1. Kezele DB, Friesen G (1993) XBT test data comparison and analysis. Sea
Technology 34: 15-22.

2. Kizu S, Hanawa K (2002) Recorder-dependent temperature error of


expendable bathythermograph. Journal of Oceanography 58: 469-476.

3. Kizu S, Hanawa K (2002) Start-up transient of XBT measurement. Deep-Sea


Res Pt I 49: 935-940.

4. Kizu S, Ito SI, Watanabe T (2005) Inter-manufacturer difference and


temperature dependency of the fall-rate of T-5 expendable bathythermograph.
Figure 8: Color photograph showing sphere at full submersion. Journal of Oceanography 61: 905-912.

5. Kizu S, Sukigara C, Hanawa K (2011) Comparison of the fall rate and


The counterpart photograph from the experiments is provided in structure of recent T-7 XBT manufactured by Sippican and TSK. Ocean
Science 7: 231-244.
Figure 7. The image is annotated to show the depth of the sphere and
the splash width. From the experiments, the splash width was found to 6. Gouretski V (2012) Using GEBCO digital bathymetry to infer depth biases in
the XBT data. Deep-Sea Res Pt I 62: 40-52.
be 0.051 m which is within approximately 6% of the simulated results.
Again, the close agreement between experiments and simulation 7. Gouretski V, Koltermann KP (2007) How much is the ocean really warming?
lends strong support to the quality of the results. Geophy Res Lett 34: L01610.

8. Gouretski VR, Reseghetti F (2010) On depth and temperature biases in


A front-lit color counterpart of Figure 7 is shown in Figure 8. This bathythermograph data: Development of a new correction scheme based on
color-version image reveals more fine detail regarding the splash size analysis of a global ocean database. Deep-Sea Res Pt I 57: 812-833.
and the formation of droplets. Droplet formation was not able to be 9. Stark J, Gorman J, Hennessey M, Reseghetti F, Willis J, et al. (2011) A
captured by the numerical simulation. computational method for determining XBT depths. Ocean Sciences 7: 733-
743.
It was shown in reference [19] and in the literature review
10. Abraham JP, Gorman JM, Reseghetti F, Sparrow EM, Minkowycz WJ (2012)
provided therein that the results shown here for the dimensionless
Turbulent and transitional modeling of drag on oceanographic measurement
drag coefficient are independent of many of the details of the devices. Computational Fluid Dynamics and its Applications 2012: 567864.
experiment or simulation. For instance, in reference [19] it was
11. Abraham JP, Gorman JM, Reseghetti F, Trenberth K, Minkowycz WJ
shown that the simulations matched experiments extremely well (2011) A new method of calculating ocean temperatures using expendable
even though there were significant differences in the dimensions bathythermographs. Energy and Environment Research 1: 2-11.

Volume 3 Issue 2 1000126 Page 6 of 4


Citation: Gorman J, Abraham J, Schwalbach D, Shepard T, Stark J, et al.,(2014) Experimental Verification of Drag Forces on Spherical Objects Entering Water.
J Mar Biol Oceanogr 3:2.

doi:http://dx.doi.org/10.4172/2324-8661.1000126

12. Abraham JP, Gorman JM, Reseghetti F, Sparrow EM, Minkowycz WJ (2012) 23. Shiffmann M, Spencer DC (1947) The flow of an ideal incompressible fluid
Drag coefficients for rotating expendable bathythermographs and the impact about a lens. Quart Appl Math 5: 270-288.
of launch parameters on depth predictions. Numer Heat Tr A-Appl 62: 25-43.
24. Nisewanger CR (1961) Experimental determination of pressure distribution on
13. Abraham JP, Baringer M, Bindoff NL, Boyer T, Cheng LJ, et al. (2013) A a sphere during water entry. NAVWEPS Report 7808, US Naval Ordinance
review of global ocean temperature observations: Implications for ocean heat Test Station China Lake CA.
content estimates and climate change. Reviews of Geophysics 51: 450-483.
25. Baldwin JL (1975) Vertical water entry of some ogives, cones, and cusps.
14. Heinmiller RH, Ebbesmeyer CC, Taft BA, Olson DB, Nikitin OP, et al. (1983) National Technical Information Services, US Department of Commerce, AD-
Systematic errors in expendable bathythermograph (XBT) profiles. Deep-Sea A009 300.
Res 30: 1185-1197.
26. Baldwin JL (1975) Prediction of vertical water-entry forces on ogives from
15. Hanawa K, Yoritaka Y (1987) Detection of systematic errors in XBT data and cone data. NSWC Technical Report, Naval Surface Weapons Center, Silver
their correction. Journal of the Oceanographic Society of Japan 43: 68-76. Spring, MD, USA.

16. Hanawa K, Rual P, Bailey R, Sy A, Szabados M (1995) A new depth time 27. Baldwin JL, Steves HK (1975) Vertical water entry of spheres, NSWC
equation for Sippican or TSK T-7, T-6 and T-4 expendable bathythermographs Technical Report, Naval Surface Weapons Center, Silver Spring, MD, USA.
(XBT). Deep-Sea Res Pt I 42: 1423-1451.
28. Moghisi M, Squire PT (1981) An experimental investigation of the initial force
17. Green AW (1984) Bulk dynamics of the expendable bathythermograph of impact on a sphere striking a liquid surface. J Fluid Mech 108: 133-146.
(XBT). Deep-Sea Res 31: 415-426.
29. Shi HH, Takami T (2001) Some progress in the study of the water entry
18. Xiao H, Zhang X (2012) Numerical investigations of the fall rate of a sea- phenomenon. Exp Fluids 30: 475-477.
monitoring probe. Ocean Eng 56: 20-27.
30. Park MS, Jung YR, Park WG (2003) Numerical study of impact force and
19. Abraham J, Gorman J, Reseghetti F, Sparrow E, Stark J, at al. (2014) ricochet behavior of high-speed water-entry bodies. Comput Fluids 32: 939-
Modeling and numerical simulation of the forces acting on a sphere during 951.
early-water entry. Ocean Eng 76: 1-9.
31. Truscott TT, Techet AH (2009) Water entry of spinning spheres. J Fluid Mech
20. Richardson EG (1948) The impact of a solid on a liquid surface. Proceedings 625: 135-165.
Physical Society 61: 352-367.
32. Zhao MH, Chen XP (2012) A combined data processing method on water
21. Shiffmann M, Spencer DC (1945) The force of impact on a sphere striking a impact force measurement. J Hydrodynamics Ser B 24: 692-701.
water surface. AMP Report 42, 1R. AMG-NYU, No.105.
33. Menter F (1994) Two-equation eddy-viscosity turbulence models for
22. Shiffmann M, Spencer DC (1945) The force of impact on a sphere striking a engineering applications. AIAA J 32: 1598-1605.
water surface. AMP Report 42, 2R. AMG-NYU, No.133.

Author Affiliations Top

Department of Mechanical Engineering, University of Minnesota,


1

Minneapolis, USA
2
School of Engineering, University of St. Thomas, St. Paul, USA
3
ENEA, UTMAR-OSS, Forte S. Teresa, Italy

Submit your next manuscript and get advantages of SciTechnol


submissions
50 Journals
21 Day rapid review process
1000 Editorial team
2 Million readers
More than 5000
Publication immediately after acceptance
Quality and quick editorial, review processing

Submit your next manuscript at www.scitechnol.com/submission

Volume 3 Issue 2 1000126 Page 7 of 4

View publication stats

You might also like