Sei sulla pagina 1di 9

Omer Kemal Kinaci1

Assistant Professor
YTU Naval Architecture and Marine
Engineering Department,
Yildiz Technical University,
Barbaros Bulvari,
Besiktas, Istanbul 34349, Turkey;
Marine Renewable Energy Laboratory,
University of Michigan,
Computational and Experimental
Ann Arbor, MI 48109
e-mail: kinaci@yildiz.edu.tr Assessment of Turbulence
Sami Lakka
Lakka Technologies, Stimulation on Flow Induced
Vatsoilantie 3,
Lemp€a€al€a 37500, Finland;
Marine Renewable Energy Laboratory,
Motion of a Circular Cylinder
University of Michigan,
Vortex-induced vibrations (VIVs) are highly nonlinear and it is hard to approach the
Ann Arbor, MI 48109
problem analytically or computationally. Experimental investigation is therefore essen-
e-mail: sami.lakka@gmail.com
tial to address the problem and reveal some physical aspects of VIV. Although computa-
tional fluid dynamics (CFDs) offers powerful methods to generate solutions, it cannot
Hai Sun replace experiments as yet. When used as a supplement to experiments, however, CFD
Assistant Professor
can be an invaluable tool to explore some underlying issues associated with such compli-
College of Aerospace and Civil Engineering,
cated flows that could otherwise be impossible or very expensive to visualize or measure
Harbin Engineering University,
experimentally. In this paper, VIVs and galloping of a cylinder with selectively distrib-
Harbin 150009, China;
uted surface roughness—termed passive turbulence control (PTC)—are investigated
Marine Renewable Energy Laboratory, experimentally and computationally. The computational approach is first validated with
University of Michigan, benchmark experiments on smooth cylinders available in the literature. Then, experi-
Ann Arbor MI 48109 ments conducted in the Marine Renewable Energy Laboratory (MRELab) of the Univer-
e-mail: hais@umich.edu sity of Michigan are replicated computationally to visualize the flow and understand the
effects of thickness and width of roughness strips placed selectively on the cylinder. The
Ethan Fassezke major outcomes of this work are: (a) Thicker PTC initiates earlier galloping but wider
Naval Architecture and Marine Engineering, PTC does not have a major impact on the response of the cylinder and (b) The amplitude
University of Michigan, response is restricted in VIV due to the dead fluid zone attached to the cylinder, which is
Ann Arbor, MI 48554 not observed in galloping. [DOI: 10.1115/1.4033637]
e-mail: ethandf@umich.edu
Keywords: vortex-induced vibrations, passive turbulence control, galloping
Michael M. Bernitsas
Professor
CVortex Hydro Energy,
Ann Arbor, MI 48109;
Marine Renewable Energy Laboratory,
University of Michigan,
Ann Arbor, MI 48109
e-mail: michaelb@umich.edu

1 Introduction cylinder motion based on the location of PTC [6]. In 2013, the
robustness of the map was tested and dominant zones were identi-
In the MRELab of the University of Michigan, flow-induced
fied [7]. Even though the PTC-to-FIM Map has become a power-
motion (FIM) is studied as a means to convert marine hydroki-
ful tool in inducing specific motions of circular cylinders, several
netic (MHK) energy to electricity using the vortex-induced vibra-
parameters remain unexplored. Experiments, albeit being the ulti-
tions for aquatic clean energy (VIVACE) energy harvester [1–5].
mate verification tool, are time consuming and hard to provide all
Turbulence stimulation by selectively distributed surface rough-
needed information. A computational tool that could predict the
ness, in the form of sand-strips, referred to as PTC, was added to
FIM of a cylinder correctly would be invaluable to study the full
oscillating cylinders in 2008 [5]. PTC enabled VIVACE to har-
parametric design space. A major side benefit of PTC was the fact
ness hydrokinetic energy from currents/tides over the entire range
that PTC enabled CFD simulations to generate results in good
of FIM including VIV and galloping. In 2011, the MRELab pro-
agreement with experiments by forcing the location of the separa-
duced experimentally the PTC-to-FIM Map defining the induced
tion point [8]. This valuable tool, along with experiments, is used
in this paper to investigate PTC design parameters such as width
1
Corresponding author. and thickness and their impact on flow features with the intent of
Contributed by the Ocean, Offshore, and Arctic Engineering Division of ASME
for publication in the JOURNAL OF OFFSHORE MECHANICS AND ARCTIC ENGINEERING.
maximizing FIM and thus, hydrokinetic energy conversion.
Manuscript received September 9, 2015; final manuscript received May 2, 2016; Passive turbulence control is a means to altering flow kinemat-
published online June 2, 2016. Assoc. Editor: Ye Li. ics around a cylinder by covering parts of the cylinder with strips

Journal of Offshore Mechanics and Arctic Engineering AUGUST 2016, Vol. 138 / 041802-1
C 2016 by ASME
Copyright V

Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


that may be smooth or have surface roughness. Roughness was
added to the whole cylinder surface before Bernitsas and Ragha-
van used selectively distributed surface roughness to enhance or
minimize VIVs, depending on the location of the strips [5]. The
strips change the boundary layer separation point and the flow
separation characteristics may be controlled passively [9,10].
Chang et al. [11] and Park et al. [6] built the PTC-to-FIM Map,
which defines zones of placement of PTC strips for enhancing or
minimizing FIM. General information on VIV is available in
review papers [12–14] and on galloping in Ref. [15]. VIVACE
converts MHK energy to mechanical in the oscillations of the cyl-
inder in FIM and subsequently converts it to electricity [3]. It uses
VIV and galloping with a single cylinder. Those motions are fur-
Fig. 1 Cylinder with PTC located at 30246 deg
ther enhanced by properly implementing gap flow with multiple
cylinders [16]. The higher and faster the FIM oscillations, the
more power the cylinder will convert from hydrokinetic to me-
chanical. Therefore, from the point of view of MHK energy har- separation point in CFD simulations is stuck at 90 deg instead of
nessing, vibration amplitude and frequency should be enhanced. oscillating around 81 deg for subcritical flow. This failure in 2D
When PTC is applied to the upstream part of the cylinder as estab- and 3D codes alike makes simulations useless even as comple-
lished in the PTC-to-FIM Map, PTC will induce galloping. PTC ments to experimental results. PTC is the only way to achieve
induces galloping instability, which has been reported to result in good agreement between experiments and CFD simulations. In
oscillation amplitude up to 3 to 4 times the diameter of the cylin- the experiments, in addition to the PTC, models reach within 2 cm
der [6,7]. of the sidewalls of the recirculating channel, thus, limiting tip-
This paper studies the effect of PTC particulars on the flow- flow effects and keeping the cylinder outside the boundary layer
induced motions of a circular cylinder, computationally and of the flow [18]. So, the 3D nature of flow kinematics in VIV is
experimentally. The physical model is described in Sec. 2. The limited in the experiments. The effect of tip flow and its interfer-
computational approach is explained in Sec. 3 and validated in ence with the 3D nature of flow in VIV are investigated in detail
Secs. 4 and 5. The PTC thickness is studied in Sec. 6, and the in Ref. [19].
PTC width in Sec. 7. CFD complements experimental results due
to the limited possibilities of visualizing the flow at these high 3 Computational Approach
Reynolds numbers (varying between 30,000 < Re < 120,000).
Flow visualization is achieved by CFD as presented in Sec. 8. Flow analysis with CFD is performed using the 2D-URANS
This facilitates discussion of the physics of the flow in VIV and approach implementing the k  x shear stress transport (SST)
galloping. Conclusions are provided in Sec. 9. turbulence model. k  xSST is a powerful model which can be
used for flows around bluff bodies. k  xSST also takes advant-
age of the strengths of the k  e turbulence model in the outer
2 Properties of the Cylinder flow region [20].
To avoid extended computational periods, the problem is simpli-
The cylinder properties used in this study both in the computa- fied by using a 2D approach. Two-dimensional flow assumption to
tional and experimental studies are provided in Table 1. Mass solve for VIV is a common practical approach implemented by
ratio, m*, is the oscillating mass over mass displaced mass. many researchers in the field. A broad literature review on compu-
The forward (upstream) stagnation point is referred to as the 0 tational approach is provided by Sarpkaya [13]. In two dimensions,
point. Positioning of the PTC on the cylinder is measured from the oscillating mass is nondimensionalized by the total length of the
the forward stagnation point. An example of PTC positioning on cylinder. Due to the 2D structure of the computational approach,
the cylinder at 3046 deg is shown in Fig. 1. some flow characteristics like cross-flows, tip vortices, and cellular
The roughness strips are placed on the cylinder symmetrically shedding are neglected. The grid structure used in this study is
with respect to the flow as shown in Fig. 1. The system has one shown in Fig. 2. A close-up view of the grid structure in the vicinity
degree-of-freedom. The in-flow movement is constrained and of the cylinder is given in Fig. 3, where the roughness strips can
thus, the cylinder is allowed to move only in the direction perpen- also be seen located symmetrically with respect to the flow. Quadri-
dicular to the flow. lateral elements are used near the cylinder and its wake, while tetra-
The PTC used in experiments and CFD in the MRELab have hedral elements are used in the outer region. The inner domain,
proven to result in agreement between CFD and experiments for which has quadrilateral elements, moves together with the cylinder
the following reason [8,9]. Alike, two-dimenisional (2D)- and is not deforming. The tetrahedral elements in the outer domain
Unsteady Reynolds-Averaged Navier-Stokes (URANS), three- are deformed and remeshed in response to cylinder movement. The
dimensional (3D)-URANS, large eddy simulation, detached eddy boundaries of the whole domain are constructed far away so that
simulation codes fail to simulate cylinder flows at Reynolds num- they do not have any effect on the cylinder’s VIV and galloping
bers greater than 10,000 [17]. The source of the error is the poor response. The error limit is set to be on the order of 105 and nu-
prediction of the motion of the flow separation point [8]. PTC in merical results converge well.
the form of straight sandpaper forces separation at the leading The time step size is chosen to follow the Courant–
edge of the sandpaper strip. Actually, for Re > 10,000, the Friedrichs–Lewy condition, which is given as

Xn
uxi
Table 1 Properties of the cylinder used in this study Dt  Cmax
i¼1
Dx i
Cylinder diameter D 0.0889 m
Cylinder length L 0.9144 m
Total oscillating mass mosc 9.784 kg where Cmax is selected to be equal to one. Here, uxi , Dxi , and Dt
Mass ratio m* 1.725 denote the velocity, height of the grid elements on the cylinder,
Natural frequency in water with inviscid added mass fn,w 1.1183 s1 and time-step size, respectively. i denotes the degree-of-freedom
Damping ratio f 0.0158 of the system, and n is the number of degrees-of-freedom, which
for the current case is n ¼ 1.

041802-2 / Vol. 138, AUGUST 2016 Transactions of the ASME

Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 2 Portion of the grid structure

Fig. 4 Validation of the computational approach with experi-


Fig. 3 Close-up view of the grid in the cylinder vicinity ments [21]

The roughness strips used in experiments in this study are com- cylinder at low Reynolds numbers [21]. These measurements
posed of backing paper with grits of different size on it. In CFD, were taken at Re ¼ 3800 and are valid till Re of about 10,000.
however, thickness is modeled as a step on the cylinder surface These numbers are in the TrSL2 flow regime. The computational
with total thickness equal to the sum of the backing paper (h) and and experimental amplitude responses of the cylinder in the
the average grit height (k) to make the complexity of grid genera- TrSL2 regime are given in Fig. 4. It may be said that the computa-
tion manageable. tional approach is successful to accurately predict pre-VIV range
About 12,000 elements are used in the whole fluid domain at and initial branch while it can only partially capture the upper
the start of the simulation. The number of elements, however, is branch, lower branch, and desynchronization.
subject to change as new elements may be formed during auto- Computationally, capturing the upper branch is a challenging
matic remeshing. Depending on the response amplitude of the issue as many researchers in the field have tried different methods
oscillating cylinder, the number of elements may reach up to or approaches to accurately calculate the range of synchronization
16,000. The computational method reaches steady-state at around and the maximum amplitude achieved. One could find many com-
10 s of real-time simulation. So unlike experiments, the maximum putational results implementing Reynolds-Averaged Navier–Stokes
point that the cylinder reaches at each oscillation does not change. Equations (RANSE) methods, large eddy simulation, or even direct
Once URANS reaches a steady-state and finds its stable ampli- numerical simulation, but none of these methods are completely
tude, the cylinder oscillates at that maximum amplitude for the accurate or computationally economical. Other computational
duration of the simulation. results implementing RANSE methods are summarized in Refs.
A real-time simulation of 10 s takes around 1.5 h of computa- [8], [22], and [23]. A comparison of results implementing 2D-
tion time with an Intel Xeon central processing unit E5-2630 at URANS is given in Ref. [24].
2.30 GHz. The workstation that is used for these simulations has a
64-bit operating system with 64 GB of installed memory (RAM).
The code uses 12 cores with two processors in each core. 5 Computationally and Experimentally
Observed FIM
The effect of the PTC thickness is investigated for a fixed loca-
4 CFD Validation With a Benchmark Case tion of the two strips on the cylinder selected to cover the surface
The computational approach is first validated with the bench- between 30 deg and 46 deg on both sides of the cylinder with
mark experiments of Khalak and Williamson for a smooth respect to the forward point of symmetry. The properties of

Journal of Offshore Mechanics and Arctic Engineering AUGUST 2016, Vol. 138 / 041802-3

Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Table 2 Properties of roughness strips

Thickness of the backing paper (h) Average grit height (k) Total height in CFD (h þ k)
3 3
P60 6.50  10  D 3.03  10  D 9.53  103  D
P180 5.94  103  D 1.41  103  D 7.35  103  D

the roughness strips (commercial sandpaper) used are listed in Upper branch: In the TrSL3 flow regime, the upper branch is
Table 2. wide with range 5 < U < 8 for a smooth cylinder and character-
Figure 5 shows the amplitude response for a smooth cylinder ized by a nearly steady increase in amplitude and frequency
and for two single cylinders with PTC. One PTC-cylinder is response leading to the lower branch. The upper branch is wider
appended with P60 commercial-grade sandpaper and the other with experimental setups that can better approximate 2D flows.
with P180. Thickness and roughness are important parameters and The gap between the tip of the cylinder and the channel wall is
are shown in Table 2 [6,7]. Figure 5 shows CFD simulations per- smaller in such setups. Thus, the flow induces lift more efficiently
formed in this study for the two PTC-cylinders in comparison to and cannot escape through the tips.
experimental measurements. PTC forces the flow separation at CFD and experiments show the same dependence on Reynolds
leading edge of the sandpaper strip and, thus, alters the flow- number. For a PTC-cylinder, the upper branch starts at
induced motion of the cylinder. Figure 6 shows the corresponding U  ¼ 4:5  5 and extends through the range of transition from
data for the frequency response of the smooth and rough VIV to galloping around U ¼ 11. Both computationally and
cylinders. experimentally, galloping occurs prior to the lower branch and
The form of the FIM response of the cylinders follows the typi- desynchronization of VIV; thus, those two VIV branches are not
cal response observed and measured consistently in the MRELab observed. For cases of higher damping, the amplitude of the VIV
for 30; 000 < Re < 120; 000, which falls in the high-lift TrSL3 branches reduces. Further, galloping initiates later, thus separating
flow regime [2,4,11]. VIV response in TrSL3 differs dramatically VIV from galloping [11]. PTC may initiate galloping but also
from TrSL2 VIV response even though each branch is clearly results in reduced VIV response.
identifiable in both flow regimes [14]. The major difference is There is an obvious bias of computational and experimental
that in TrSL3, the upper branch overtakes the lower branch and results at this branch both with smooth and rough cylinders. Com-
amplitude increases linearly rather than remaining constant [2,4]. putationally, the amplitude response increases linearly in the
On the basis of Figs. 5 and 6, we can make the following upper branch for a PTC-cylinder while it is nearly constant experi-
observations. mentally. The oscillation frequency, on the other hand, is higher
Pre-VIV range: For U < 4; a smooth cylinder does not as compared to computational results. CFD cannot capture the
exhibit FIM. For PTC-cylinders, the pre-VIV range starts at flow physics of the late upper branch and therefore it may be said
U  ¼ 4. Pre-VIV range occurs slightly earlier in CFD generated that computational results are not in good accordance with experi-
results and it starts at around U  ¼ 3 both with smooth and rough ments at this flow range of VIV for the PTC-cylinder. For the
cylinders. smooth cylinder case, the large amplitudes occurring in this phase
Initial branch: It is short characterized by a large jump in ampli- of flow were not captured computationally. This is a typical error
tude and frequency response. For a smooth cylinder, it is in encountered in many computational approaches. A discussion was
the range 4 < U < 5. For the PTC-cylinder, it is in the range held relevant to this in Ref. [24] and in the same article, a compar-
4 < U  < 4:5 experimentally and 3 < U  < 4 computationally. ison of computational results for the benchmark experiments

Fig. 5 Amplitude response of smooth cylinder versus PTC- Fig. 6 Frequency response of smooth cylinder versus PTC-
cylinders with variable PTC thickness cylinders with variable PTC thickness

041802-4 / Vol. 138, AUGUST 2016 Transactions of the ASME

Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


conducted in Ref. [21] for flows in the TrSL2 regime are given. 7 Effect of the PTC Width
For flows at higher Reynolds numbers like in the TrSL3 re-
The effect of the PTC width is experimentally and computa-
gime, similar problems were observed in this study and also in
tionally investigated in this section by comparing the results of
Ref. [24].
the 16 deg PTC P60 in Sec. 6 with 8 deg PTC P60 strips. The start-
Lower branch: For smooth cylinders at this flow regime
ing points of the roughness strips are set at 30 deg. The narrower
(TrSL3), the lower branch is not present. This significant VIV
strip (8 deg) ends at 38 deg while the broader one (16 deg) ends
response as compared to TrSL2 flow regime is first discussed in
at 46 deg. The amplitude and frequency responses for 8 deg of
Ref. [4]. For PTC-cylinders, the lower branch is not present. The
coverage and their comparison with 16 deg of coverage are given
beginning of the lower branch is overtaken by the upper branch
in Figs. 7 and 8, respectively. The following observations can be
while the end of it is overtaken by the onset of galloping. Although
made:
the form of the curve generated by CFD does not capture the ampli-
VIV region: As shown in Fig. 7, the PTC width does not have a
tude and frequency responses of the cylinder quite well, it may be
major impact on the oscillation amplitude and frequency on VIV.
said that the lower branch is not present in CFD calculations as
well.
VIV to galloping transition: Smooth cylinders do not go into
galloping unless there is an asymmetry introduced by geometry or
upstream nonuniform flow. PTC-cylinders go into galloping as
represented by the high amplitudes after U  ¼ 12. In the range
11 < U  < 12, both mechanisms of VIV and galloping work
together to induce sudden amplitude increase and frequency
change. After certain amplitude is reached, the VIV mechanism
desynchronizes and galloping takes over. Transition from VIV to
galloping is more gradual computationally than experimentally.
Nevertheless, the characteristic amplitude increase in transition
from VIV to galloping is shown by CFD as well. Using CFD flow
visualization in Sec. 5, the underlying hydrodynamic mechanism
is also demonstrated. At this point, however, it should still be
highlighted that computationally, the transition to galloping is not
as sharp as it is in the experiments. Onset of galloping cannot be
pinpointed from computational results.
Fully developed galloping: For 12:5 < U  , galloping is fully
developed and the amplitude of oscillation reaches values that
result in more complex vortex structures that do not synchronize
with the galloping motion. The galloping amplitude continues to
increase albeit slower than shown experimentally. The underlying
hydrodynamic mechanism is captured by CFD as discussed in
Sec. 5 and is consistent with experimental measurements by
Park et al. [6].

6 Effect of the PTC Thickness Fig. 7 Effect of PTC coverage in amplitude response
Based on the computational and experimental results in Sec. 5,
we can make the following observations regarding the effect of
PTC thickness on FIM.
VIV amplitude: As shown in Fig. 5, the amplitude response of
the cylinder remains almost steady and PTC thickness has only a
slight effect in the upper branch in VIV.
VIV frequency: As shown in Fig. 6, frequency response
increases with PTC thickness throughout the entire VIV range
considered in this work.
Galloping onset: PTC thickness affects the onset of galloping.
Thicker PTC induces galloping slightly earlier than the thinner
one. The thickness affects the point of flow separation and possi-
ble reattachment. It has been reported by Chang et al. that the
thicker PTC promotes earlier galloping [11]. The same applies
here where the thicker PTC (P60) induces galloping earlier than
the thinner one (P180).The PTC thickness has negligible effect in
the galloping region, where A has the same form for different
values of the PTC thickness. Galloping is triggered by geometric
asymmetry, resulting in negative damping and thus instability.
The instability is slightly stronger for thicker PTC, which is con-
sistent with the conclusions by Blevins [15].
Galloping amplitude: The PTC thickness does not affect the
amplitude response significantly in the galloping region.
Galloping frequency: The thicker PTC (P60) causes f* ¼ fosc/
fn,water to have smaller values throughout the whole VIV range.
The shape of the f  versus U* curve does not change because gal-
loping is an instability. The PTC thickness affects the magnitude
of f*. When compared to the smooth cylinder, PTC increases the
frequency of the cylinder. Fig. 8 Effect of PTC coverage in frequency response

Journal of Offshore Mechanics and Arctic Engineering AUGUST 2016, Vol. 138 / 041802-5

Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


PTC helps alter the point of flow separation from its smooth cylin- of the cylinder. This is also captured computationally, although it
der location. Its width (coverage) plays a secondary role in VIV as is found out that the narrower PTC induces slightly higher
long as the entire sand strip remains inside the same zone of the frequency in the VIV range.
PTC-to-FIM Map [6,7]. Additional CFD results with relevant flow visualization are pro-
Galloping region: Narrower PTC generates lower amplitudes in vided in Sec. 8.
the galloping range but has negligible effect in the VIV range.
Similarly, Chang et al. [11] observed that the bigger PTC cover-
age leads to increased amplitudes in the galloping range. They 8 Computational Visualization
also observed that the bigger PTC coverage leads to decreased Upon verification and experimental validation of the CFD code
amplitudes in the VIV range, which is consistent with the PTC-to- applied to a specific problem, the code can become a very useful
FIM Map, as their PTC was closer to the dominant suppression tool in investigating the FIM under study. To gain a greater insight
zone [6]. Their PTC coverage started at 40 deg. When the robust- in the flow details, the flow has to be visualized. It is difficult to
ness of the PTC-to-FIM map was studied [6], it was concluded visualize the flow experimentally at high Reynolds numbers using
that the strong suppression zone is the dominant zone when PTC particle image velocimetry. Therefore, to visualize the flow char-
covers multiple zones. That is, the conclusion in this paper is con- acteristics in VIV and galloping, CFD can provide a unique tool.
sistent with Ref. [11]. In this section, pressure distributions as in Fig. 9 and shear
Similar results were observed regarding the effect of PTC thick- stress around the cylinder are used to integrate pressure and
ness using a computational approach. For some part of the gallop- calculate force. Velocity vector fields in the body vicinity in
ing range, CFD returns higher amplitudes for the narrower PTC Figs. 10–13 are used to visualize separation of the flow. The fol-
although the difference is really small and hard to distinguish in lowing observations can be made based on Figs. 9–13.
Fig. 7. However, it is hard to generalize this result as this is only Amplitude response: The difference in amplitudes can be con-
valid for two data points in the range covered in this study. More firmed by the pressure differences between the upper and lower
data points are needed in the galloping phase to establish a solid parts of the cylinder as well. The pressure coefficients at the low-
conclusion. est position of the cylinder for reduced velocity U  ¼ 14 for the
Frequency response: It is shown in Fig. 8. It is observed that the cylinders with PTC P60 and P180 are given in Fig. 9. The thicker
PTC thickness has a negligible effect on the frequency response PTC generates higher amplitudes experimentally and

Fig. 9 Four quadrants of the cylinder (left). Pressure coefficient distribution along the
cylinder with PTC P60 and P180 at U  5 14 (right).

Fig. 10 Velocity vectors at U  5 7 for PTC P60 2 16 deg coverage (left). Zoomed in version
(right).

041802-6 / Vol. 138, AUGUST 2016 Transactions of the ASME

Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 11 Velocity vectors at U  5 14 for PTC P60 2 16 deg coverage (left). Zoomed in version
(right).

Fig. 12 Velocity vectors at U  5 7 for PTC P60 2 8 deg coverage (left). Zoomed in version
(right).

Fig. 13 Velocity vectors at U  5 14 for PTC P60 2 8 deg coverage (left). Zoomed in version
(right).

computationally (see Fig. 5) in the transition from VIV to gallop- Separation: Figures 10–13 present the velocity vectors
ing region. It also generates higher amplitudes in the galloping around the cylinder for 16 deg and 8 deg coverage at U  ¼ 7
region but with smaller differences as the vortex shedding mecha- and U  ¼ 14. These figures capture the flow speed and vector
nism acts out of synchronization with the galloping instability field at the lowest position of the cylinder oscillation. The
driving mechanism. boundary layer separation on the lower side of the cylinder
Vortex shedding: The cylinder flow is studied in four quadrants occurs earlier at U  ¼ 7 when compared to the separation at
as shown in Fig. 9. It may be noticed directly that there is a higher U  ¼ 14. There is a reattachment of the flow at the upper part
pressure gradient at the rear part of the cylinder (between quad- of the cylinder, which is not observed at the lower part inde-
rants II and III), which is where vortex shedding occurs. Higher pendent of the flow speed. The formation of the flow charac-
pressure gradients lead to stronger vortex shedding which, in turn, teristics shown in Figs. 10–13 is periodic and at the upper end
results in higher amplitudes. The protrusions in the pressure coef- of the oscillation, they are approximately reversed. The bound-
ficient graph are the places of the roughness strips. ary layer separation and flow reattachment at the upper part of

Journal of Offshore Mechanics and Arctic Engineering AUGUST 2016, Vol. 138 / 041802-7

Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


the cylinder occur later in galloping compared to the VIV adopted in this paper was insufficient at some specific phases of
range. fluid induced motions like the VIV phase of the cylinder. In
Trapped fluid: The trapped dead fluid in the VIV region future studies, denser grids in the near vicinity of the cylinder
(U  ¼ 7) is greater when compared to the galloping region and particularly the PTC location will be implemented. The
(U  ¼ 14). The higher flow momentum in galloping, compared to complementary information that CFD visualization contributes
VIV, pushes the trapped dead fluid zone to a downstream part of is most valuable in interpreting numerical and experimental
the cylinder. In addition, recall that in galloping, vortex shedding information based on the location of separation and reattach-
is not the driving mechanism. On the contrary, vortex shedding is ment of the flow, which is difficult to visualize in the high
desynchronized and may act against the driving force in galloping. Reynolds number range of this study.
In the VIV region, the trapped dead fluid restricts the amplitude of
the cylinder when compared to the cylinder in the galloping
region. In the galloping region, the dead fluid zone is not carried
with the cylinder in oscillation. At U  ¼ 7, which is in the VIV Acknowledgment
region, the vortex that will be shed once the cylinder reaches its
The financial support of TUBITAK for the first author is grate-
highest position is built at the upper part of the cylinder and wait-
fully acknowledged. This paper was prepared under Cooperative
ing to be shed. At U  ¼ 14, which is in the galloping region, the
Agreement No. DE-EE0006780 between Vortex Hydro Energy,
vortex formed at the upper part of the cylinder is about to be shed.
Inc., and the U.S. Department of Energy. The Marine Renewable
In the galloping region, the number of shed vortices per half cycle
Energy Laboratory is a subcontractor through the University of
is higher when compared to the VIV region. The trapped dead
Michigan.
fluid in Figs. 10 and 12 increases the added mass at U ¼ 7. This
CFD result is consistent with the added mass curve in Ref. [25],
showing that the added mass decreases with respect to increasing
fluid velocity [25].
References
[1] Bernitsas, M. M., Raghavan, K., Ben-Simon, Y., and Garcia, E. M. H., 2008,
“VIVACE (Vortex Induced Vibration Aquatic Clean Energy): A New Concept
in Generation of Clean and Renewable Energy From Fluid Flow,” ASME J.
9 Conclusions Offshore Mech. Arct. Eng., 130(4), p. 041101.
Selectively distributed local roughness was applied to the sur- [2] Bernitsas, M. M., Ben-Simon, Y., Raghavan, K., and Garcia, E. M. H., 2009,
“The VIVACE Converter: Model Tests at Reynolds Numbers Around 105,”
face of an otherwise smooth circular cylinder in the form of sand ASME J. Offshore Mech. Arct. Eng., 131(1), p. 011102.
strips to control passively the flow kinematics around the cylinder [3] Bernitsas, M. M., and Raghavan, K., 2009, “Fluid Motion Energy Converter,”
in FIMs in a steady uniform flow. This method of altering the flow U.S. Patent and Trademark Office, U.S. Patent No. 7,493,759 B2.
is called PTC. Application of PTC has been shown in previous [4] Raghavan, K., and Bernitsas, M. M., 2011, “Experimental Investigation of
Reynolds Number Effect on Vortex Induced Vibration of Rigid Cylinder on
work to result in good agreement between CFD and experiments Elastic Supports,” Ocean Eng., 38(5–6), pp. 719–731.
even in the TrSL3 flow regime. On the contrary, CFD results for [5] Bernitsas, M. M., and Raghavan, K., 2011, “Enhancement of Vortex Induced
smooth cylinders agree with experiments for lower Reynolds Forces and Motion Through Surface Roughness Control,” U.S. Patent and
numbers only. Nevertheless, CFD results in the TrSL3 flow Trademark Office, U.S. Patent No. 8,042,232 B2.
[6] Park, H. R., Kumar, R. A., and Bernitsas, M. M., 2013, “Enhancement of
regime can be used for comparative study of cases. Some of the Flow Induced Motions of Rigid Circular Cylinder on Springs by Localized
effects of PTC are studied as function of PTC parameters compu- Surface Roughness at 3  104  Re  1.2  105,” Ocean Eng., 72(1),
tationally and are compared to experiments. Computational results pp. 403–415.
were validated quantitatively in comparison to experimental [7] Park, H. R., Bernitsas, M. M., and Chang, C. C., 2013, “Robustness of the Map
of Passive Turbulence Control to Flow-Induced Motions for a Circular Cylinder
results. Then, CFD visualization of the flow was used to study at 30,000 < Re < 120,000,” 31st OMAE 2013 Conference, Nantes, France, June
flow properties that would be difficult to study experimentally. 9–14, Paper No. 10123.
The major conclusions of this paper are as follows: [8] Wu, W., Bernitsas, M. M., and Maki, K., 2014, “RANS Simulation Versus
Experiments of Flow Induced Motion of Circular Cylinder With Passive Turbu-
 CFD visualization enables studying flow kinematic features lence Control at 35,000 < Re < 130,000,” ASME J. Offshore Mech. Arct. Eng.,
at the boundary layer scale, such as flow separation and reat- 136(4), p. 041802.
tachment, which would be difficult to visualize experimen- [9] Ding, L., Bernitsas, M. M., and Kim, E. S., 2013, “2D URANS vs. Experiments
of Flow Induced Motions of Two Circular Cylinders in Tandem With Passive
tally in the TrSL3 flow regime. Turbulence Control,” Ocean Eng., 72(2), pp. 429–440.
 With PTC, the amplitude response of the cylinder remains [10] Ding, L., Zhang, L., Kim, E. S., and Bernitsas, M. M., 2015, “2D-
steady in the VIV range. URANS vs. Experiments of Flow Induced Motions of Multiple Circular
 Galloping is initiated earlier with a thicker PTC. Cylinders With Passive Turbulence Control,” J. Fluids Struct., 54,
pp. 612–628.
 The onset of galloping using CFD with the k  x SST turbu- [11] Chang, C. C., Kumar, R. A., and Bernitsas, M. M., 2011, “VIV and Galloping
lence model is at the same location as shown in experiments, of Single Circular Cylinder With Surface Roughness at 3.0  104  Re 
albeit not as steep. 1.2  105,” Ocean Eng., 38(16), pp. 1713–1732.
 CFD results reveal that the width of the PTC does not have a [12] Govardhan, R., and Williamson, C. H. K., 2000, “Modes of Vortex Formation
and Frequency Response of a Freely Vibrating Cylinder,” J. Fluid Mech., 420,
major impact on the amplitude and the frequency response of pp. 85–130.
the cylinder. This is in agreement with the experimental con- [13] Sarpkaya, T., 2004, “A Critical Review of the Intrinsic Nature of Vortex-
clusion in Ref. [6] as long as the roughness strip remains in Induced Vibrations,” J. Fluids Struct., 19(4), pp. 389–447.
one zone of the PTC-to-FIM Map [6,7]. [14] Williamson, C. H. K., and Govardhan, R., 2004, “Vortex-Induced Vibrations,”
Annu. Rev. Fluid Mech., 36(1), pp. 413–455.
 CFD results reveal the difference between near-cylinder flow [15] Blevins, R. D., 2001, Flow-Induced Vibration, 2nd ed., Van Nostrand Reinhold,
characteristics in the VIV and galloping phases. The dead New York.
fluid zone attached to the cylinder in the VIV region restricts [16] Kim, E. S., Bernitsas, M. M., and Kumar, A. R., 2011, “Multi-Cylinder Flow
the maximum amplitude achieved. The dead fluid zone is at Induced Motions: Enhancement by Passive Turbulence Control at
28,000 < Re < 120,000,” ASME J. Offshore Mech. Arct. Eng., 135(1),
a later part of the cylinder in the galloping phase and it p. 021802.
breaks loose quicker when compared to the VIV phase. The [17] Wanderley, J. B. V., Sphaier, S. H., and Levi, C., 2009, “A Numerical Investi-
dead fluid zone is not carried with the cylinder during oscilla- gation of Vortex Induced Vibration on an Elastically Mounted Rigid Cylinder,”
tion, which increases the maximum amplitude in the gallop- 28th OMAE’09 Conference, Paper No. 79725.
[18] Bernitsas, M. M., 2016, “Harvesting Energy by Flow Included Motions,”
ing phase. Springer Handbook of Ocean Engineering, M. R. Dhanak and N. I. Xiros, ed.,
Springer-Verlag, Berlin/Heidelberg, Germany, Chap. 47.
Future studies using CFD will cover broader ranges of PTC [19] Kinaci, O. K., Lakka, S., Sun, H., and Bernitsas, M. M., 2016, “Effect of Tip-
width and location to compare to experimental results that Flow on Vortex Induced Vibration of Circular Cylinders for Re < 1.2*105,”
defined the PTC-to-FIM map [6,7]. The numerical methodology Ocean Eng., 117, pp. 130–142.

041802-8 / Vol. 138, AUGUST 2016 Transactions of the ASME

Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


[20] Menter, F. R., 1994, “Two-Equation Eddy-Viscosity Turbulence Models for [23] Pan, Z., Cui, W., and Miao, Q., 2007, “Numerical Simulation of Vortex-
Engineering Applications,” AIAA J., 32(8), pp. 1598–1605. Induced Vibration of a Circular Cylinder at Low Mass-Damping Using RANS
[21] Khalak, A., and Williamson, C. H. K., 1996, “Dynamics of a Hydroelastic Code,” J. Fluids Struct., 23(1), pp. 23–37.
Cylinder With Very Low Mass and Damping,” J. Fluids Struct., 10(5), [24] Kinaci, O. K., 2016, “2D URANS Simulations of Vortex Induced Vibrations of Cir-
pp. 455–472. cular Cylinder at TrSL3,” J. Appl. Fluid Mech., 9(4) (published online).
[22] Guilmineau, E., and Queutey, P., 2002, “A Numerical Simulation of Vortex [25] Vikestad, K., Vandiver, J. K., and Larsen, C. M., 2000, “Added Mass and
Shedding From an Oscillating Circular Cylinder,” J. Fluids Struct., 16(6), Oscillation Frequency for a Circular Cylinder Subjected to Vortex-Induced
pp. 773–794. Vibrations and External Disturbance,” J. Fluids Struct., 14(7), pp. 1071–1088.

Journal of Offshore Mechanics and Arctic Engineering AUGUST 2016, Vol. 138 / 041802-9

Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Potrebbero piacerti anche