You are on page 1of 20

Bio-Inspired Passive Drag Reduction Techniques: A Review

Hayder A. Abdulbari[1,2],*, Hassan D. Mahammed[1], Zulkefli B. Y. Hassan[1]

Abstract
It was believed that fluid flow and the laminar to
turbulent transition delay were more easily controlled on smooth surfaces until the discovery of the
grooved shark skin surface that changed the whole
idea of how smooth the surface should be to have
high flow in submerged surfaces. Riblets have
gained renewed interest in academic fields of study
and in industry due to several advantages in manip-

ulating the turbulence boundary layer. Drag reduction using small, longitudinally grooved surface provides up to 10 % lower energy consumption in several
applications. This review provides an overview of the
mechanism of drag reduction with riblets, the different geometries and types, and the latest developments in drag reduction riblet technology.

Keywords: Drag reduction, Geometry, Riblets, Skin friction


Received: December 07, 2014; revised: January 24, 2015; accepted: January 29, 2015
DOI: 10.1002/cben.201400033

Introduction

In fluid dynamics applications, energy is used to either maintain the motion of a fluid over a solid surface or to move a solid
body through a fluid. In both cases, a substantial amount of
energy is expended to overcome the drag force. Since the energy crisis in 1973, increased interest in energy conservation has
spurred researchers to seek alternative means to reduce energy
consumption. Several studies have been conducted to find solutions for enhancing drag reduction for aeroplanes and vehicles,
transport pipelines, and other industrial applications [1].
Since the discovery of the drag reduction phenomenon, researchers have sought to decrease the effect of the drag force to
minimize its impact on the environment and economy. Wall
drag force was first explained in 1904 by Prandtl [2] when he
proved that all friction losses for fluid flow occur within a thin
layer adjacent to a solid boundary. The flow outside this layer
can be considered frictionless, and the velocity near the boundary is affected only by boundary shear. The boundary layer
generates a drag near the wall as a result of the viscous interaction between the surface and the fluid [25]. This drag is typically referred to as skin friction. The drag force is a component
of the resultant force exerted by the fluid over a body and is
oriented toward the comparative motion of the fluid.
Skin friction drag represents a significant component of the
total drag force. It comprises nearly 50 % of the total drag force
on commercial aircraft, 90 % of the total drag force on submarines, and 100 % of the total drag force in long distance pipelines
[68].
Pipeline transportation is one of the most convenient methods of conveying fluids due to its safety and cost efficiency.
Elongated pipeline systems provide the long-distance transport
of water, natural gas, crude oil, and other materials [912].

www.ChemBioEngRev.de

As a result of turbulent flow, a pressure drop is experienced


during fluid transport in a pipeline. Flow in the turbulent
regime is exemplified by fluctuations in acceleration, pressure,
and shear stress, which are all functions of time and position.
These fluctuations result in unpredictable convective flow
paths, unsteady vortices and eddies and increased skin friction.
Because of the increased drag force, the energy required to
overcome the viscous drag is spent pumping the fluid to rebuild the pressure head. In general, relative to flow in the laminar regime, turbulent fluid flow increases the cost of the system
[1317].
A number of studies have investigated various methods to
reduce losses from the skin friction drag force. Several recommendations and strategies have been adopted, but the most
commonly applied technique involves the use of minor quantities of drag reducing agents (DRA), i.e., viscous elastic chemical
substance, injected into pipelines to reduce drag reduction.
DRA, or flow improvers, have long been used in pipelines to
maximize the flow potential by increasing operational flexibility. DRA manipulates fluid behavior at the boundary layer by
modifying the characteristics of the fluid to minimize the onset
of turbulence. It was discovered accidentally in 1948 by Toms
[18] that the addition of a few ppm of a higher molecular

[1]

Prof. Hayder A. Abdulbari (corresponding author), Dr. Hassan


D. Mahammad, Prof. Zulkefli B. Y. Hassan
Faculty of Chemical and Natural Resources Engineering, University
Malaysia Pahang, Gambang 26300, Kuantan, Pahang, Malaysia
E-Mail: hayder.bari@gmail.com

[2]

Prof. Hayder A. Abdulbari


Center of Excellence for Advanced Research in Fluid Flow, University Malaysia Pahang, Gambang 26300, Kuantan, Pahang, Malaysia

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

These are not the final page numbers! &&

weight polymer into a turbulent flow could significantly reduce


the skin friction. DRAs can be classified into three major categories: long chain molecules (polymers), surface active agents
(surfactants) and injected micro bubbles, solid particles or
fibers (suspended solids) [1, 3, 19, 20].

Foreign Substance Drag Reduction


Techniques (Active Means)

Turbulent skin friction drag can be reduced by introducing various foreign substances. These include long chain polymers,
surfactants, micro bubbles, powders or tiny solid particles, and
fibers.
Numerous drag reduction studies have been conducted using
polymers as an additive to enhance the flow in pipelines. All
have concluded that minor concentrations of long chain polymers engender drag reductions of over 70 % [1, 2123]. Additionally, several other variables, such as the polymer chain
length, particle size, injection technique, and polymer interaction within the turbulent flow, were considered [5, 2429].
Despite their apparent advantage in reducing drag force, polymer-based DRAs cannot be used in a variety of applications.
For example, in the pharmaceutical and food and beverage
industries, polymer-based DRAs may alter the original characteristics of the fluids [30]. Furthermore, these polymer DRAs
lose their effectiveness with time due to rapid degradation
kinetics at high Reynolds numbers [31, 32].
Surfactant-based DRAs modify the surface properties of
liquids or solids. The majority of surfactants appear to reduce
the turbulent skin friction drag only in the presence of electrolytes. The highest percentage of drag reduction achieved by
Cho et al. [33] was 80 % in a solution containing between 1000
and 2000 ppm at 70 C of stearyl amine oxide and betaine. Surfactant-based DRAs are stable against physical decomposition,
which is an advantage over polymer-based DRAs that break
down when the strain rate is sufficiently high [5, 3337].
Fiber suspensions have long been recognized (earlier than
polymers) as drag reducing agents and were first described
by a small circle of engineers working with paper pulps. The
effectiveness of these fiber suspensions depends on its properties, such as density and concentration. Drag reduction was
found to be possible when the concentration is high enough
for fiber-fiber interactions to occur, but lower than a critical
concentration above which the suspension viscosity is dramatically increased [38]. Early observations of suspensionbased DRAs used suspensions of natural products, such as
sediments and wood fiber. The developments of suspensionbased DRAs were motivated by the need to provide accurate
hydraulic transport criteria. Early attempts to establish the
systematic effects of solids concentration, specific gravity and
duct dimensions were unsuccessful, quite possibly because
the suspended particles were not uniform and did not have
reproducible dimensions and surface texture. Several factors
affect the effectiveness of suspension-based DRAs, including
the type of internal flows (horizontal or vertical), type of fluid (liquid or gas) and the type and size of fibrous or nonfibrous particles [20, 39, 40]. The concentration where the

www.ChemBioEngRev.de

maximum drag reduction could be achieved using suspension-based DRAs is 4050 % [20, 41].

Non-Additive Drag Reduction


Techniques (Passive Means)

In the past decade, comprehensive improvements to the understanding and perception of turbulent boundary layers have
been made. Significant efforts have been devoted to find cheaper and environmentally-friendly alternative methods to reduce
the effects of drag force. Extensive reviews can be found in
Wilkinson et al. [42], Gad-el-Hak [1, 19], Walsh [43], Savill
[44], Raupach et al. [45], Meier et al. [46], Bushnell [47], Fish
and Lauder [48], and Paul et al. [49]. These studies focused on
turbulent drag reduction by passive means. Many techniques
have been proposed that involve mechanically altering flow
control surfaces. The main concept behind this technique is the
manipulation of the boundary layer by energy transfer which is
affected by the natural interaction dynamics of the fluid along
the solid boundary. In other words, the skin friction drag is
simply reduced by self-noise, applying these manipulators with
certain dimensions into the surface will shift away the turbulence structures from the wall, so the drag force will be reduced
by turbulence flow noise. These drag reduction methods do not
involve the use of additives. The passive means can be classified
into two groups based on the local position of the manipulators
relative to the turbulent boundary layer. (i) External layer
manipulators, or outer layer manipulations, where thin plates
are introduced into the external part of the flow, do not efficiently reduce the total drag in the turbulent boundary layers.
A few devices that are categorized as external layer manipulators include outer layer devices (OLDs), boundary layer devices
(BLADEs), large eddy breakup devices (LEBUs) and tandem
arrayed parallel plate manipulators (TAPPMs). (ii) Internal
layer manipulators, or thinner layer manipulators, restructure
the surfaces by altering the wall geometry in the form of small
stream-oriented striations, such as compliant walls, oscillating
walls, dimples and riblets [1, 20, 5053].
Several factors determine the percentage of drag reduction in
both types of manipulators. For the outer layer devices, these
factors include the angle of attack, the chord lengths of the
aerofoil or flat plate devices, the boundary height, the spacing
between tandem configurations, the thickness of the flat plate
device and the thickness of the boundary layer at the leading
edge of the device. For the inner layer manipulators, the height
and spacing between geometric striations are the only factors
that need to be considered in determining the percentage of
drag reduction [20]. Fig. 1 shows two schematics highlighting
these factors.
Most outer layer manipulators follow the same mechanism
as LEBUs, a popular drag reducing mechanism in the 1980s.
However, as shown by Wilkinson et al., LEBUs were not very
efficient at obtaining a definite total drag reduction in turbulent
boundary layers [42]. These devices were designed to break up
large vortices in turbulent flow and inhibit high Reynolds shear
stress near the wall. They consisted of metallic ribbons oriented
in the direction of flow. LEBUs developed substantial down-

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

These are not the final page numbers! &&

a)

b)
Figure 1. Parameters that control the percentage of drag reduction in (a) outer layer manipulators and (b) inner layer manipulators [20].

stream regions of reduced skin friction effects and artificial


boundary layers. They reduced the characteristic turbulence
scale by altering the momentum transferred from the external
flow toward the wall. Particular arrangements of the ribbons
may reduce the number of turbulent bursts and, hence, the surface friction stress. The reductions provided by these devices
net between 5 and l0% drag reductions [42, 5456]. However
Sahlin et al. [57] and Lynn et al. [58] depicted that the average
skin friction downstream of the devices decreased by 8 %. Findings of the study showed albeit the device drag was low, there
was an increase of 3 % in net drag. The study further added
that a decrease in net drag at high Reynolds numbers seems
implausible. Meanwhile, when the Reynolds number increases,
the effectiveness of LEBU decreases and it is all because of the
decrease in size of the most energetic turbulent scales that
relates to the boundary-layer thickness. This suggests that the
variance in both outer and inner scales disables the LEBU
mechanism.
In a numerical simulation study conducted by Spalart et al.
[59], they modified a turbulent flow and weakened the wall
pressure fluctuations on an airliner windshield. The results
show that the turbulent flow lost most of its sizable voluminous
bulges and generally created more minuscule outer-layer
eddies, which were harmless relative to the original immensely
colossal eddies. Spalart et al. succeeded in reducing the wall
fluctuations by simulating LEBUs. However, this was achieved
only over approximately 6 boundary layer thicknesses, when
the actual effect would need to be sustained over approximately
25 boundary-layer thicknesses. Eckhardt et al. [60] investigated
the possibility of reducing skin friction through the use of flowaligned vertical LEBUs in which LEBUs were mounted perpendicularly to the flat plate surface. The experimental results
showed a local drag reduction of approximately 2728 %.
Moreover, the vertical LEBUs were more effective in reducing
drag compared with the LEBUs directed parallel to the wall.
As evidenced by many studies, these devices produce downstream reduced skin frictions of up to 30 %. However, not all
LEBUs have a net drag reduction due to device-added drag.
This finding indicates that LEBUs have no drag reducing ability in channel flow [61].

www.ChemBioEngRev.de

A great deal of research has focused on compliant coating


techniques for drag reduction. The very first study was performed by the biologist Gray [62] in 1936, wherein he suggested that the malleable skin of a dolphin could have some
sort of damping influence on turbulence within the boundary
layer, therefore rendering fluid flow more laminar with less
friction. However, Kramer [63] was the first to study the use of
compliant coatings as a possible technique for drag reduction;
Grays idea inspired him to use elastic walls to simulate the
drag repelling nature of dolphin skin. The maximum drag
reduction found by Kramer was 60 %, which was impressive
considering that since then, the many studies using compliant
walls to reduce the drag force have never achieved that percentage [30, 6368].
A compliant wall provides a pressure field that tends to
inhibit the turbulent burst phenomenon [69]. According to
Kline et al. [70], this sequence of events consists of three
phases: lift-up, oscillation and break-up of low speed streaks
(i.e., bursts). It is generally assumed that the bursting process
also plays important roles in momentum, heat, and mass transfers. In the literature, compliant wall studies have primarily
focused on delaying the laminar to turbulence transition in
applications for marine vehicles. Compliant wall studies were
rarely investigated for other applications such as pipelines, due
to the different mechanism for drag reduction [64, 67, 68, 71].
Carpenter and Garrad [72, 73] noticed that Kramers compliant walls were hypersensitive to pressure gradients and performed quite differently under various experimental flow conditions. This outcome proved that the reported 60 % drag
reduction was credible and different experimental conditions
would be plausible explanations for why other researchers
failed to achieve 60 %.
Even though compliant walls are beneficial for marine vehicles, it is not applicable for drag reduction in pipes in which
the flow was already turbulent. However, compliant surfaces
are known to have the ability to modify turbulent flow and skin
friction drag and avoid boundary layer separations [64, 74].
Dimples are used in many heat transfer applications due to
the high contact area provided by the structures. Generally,
dimples belong to the same family as riblets in that both are
modifications superposed on smooth solid walls; however, they
are different from a design point of view. Most of the research
work cited in the literature on dimples has focused on heat
transfer [7584]. Drag reduction was also addressed in these
multi-objective studies [8591]. The early findings showed that
improvements in the heat transfer coefficient also increase
drag. Additionally, the heat transfer performance of the dimples was not always as efficient as that of other techniques, such
as fins and turbulators [7881]. The numbers of studies that
focused on the effect of dimples is very limited and contradictory. A few studies claimed drag reduction was achievable while
others found that drag increased with the addition of dimples
[7881]. On the one hand, Rohlfs et al. [92] introduced an
experimental and numerical investigation of the turbulent flow
over dimpled surfaces. They observed that the dimples had no
significant effect on the drag reduction (positive or negative)
which led them to conclude that restructuring the surfaces in a
dimple form will have a significant heat transfer effect and no
drag reduction performance. On the other hand, Yu et al. [93]

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

These are not the final page numbers! &&

reported an 18.0 % drag reduction when using dimpled surfaces (i.e., pin fin-dimple hybrid structures) and Zhong et al. [94]
reported a 15.1 % drag reduction with direct numerical simulations. The reason for this disagreement is the vast difference
between initial fluid conditions, dimple geometry, measurement conditions and evaluation methods. However, there is a
general agreement that drag reduction is achievable with
dimples despite the fact that its heat transfer performance is
better and more obvious. Lienhart et al. [95] introduced a numerical and experimental investigation of the drag reduction
using dimpled surfaces using a channel test rig. They studied
the effect of the dimples design and arrangement on the drag
reduction performance as shown in Fig. 2. They concluded that
using shallow dimples can have good or acceptable improvement in the heat transfer performance without any increase in
the pressure drop. In another words; they concluded that the
dimples designs investigated are not effective when it comes to
the drag reduction and when compared to the heat transfer
performance.

Oscillating walls are generally used in channels wherein one


or two walls undergo forced spanwise or streamwise oscillation
driven by a pump. Generally, the oscillation movement of the
walls has been shown to be effective in suppressing the turbulent structures of the fluid passing through the channels. The
wall oscillation frequency and amplitude are directly related to
its drag reduction performance; there is an expected one point
effect for each flow rate or degree of turbulence [96100]. Generally, the main mechanism controlling the drag reduction
effect using oscillating walls involves altering the boundary
layer conditions and creating order in a previously chaotic, turbulent flow. Choi and Clayton [101] conducted very important
research work on the drag reduction using spanwise oscillating
walls. The study was performed in an open-return, low-speed
wind tunnel. They found that 45 % drag reduction can be
achieved through their design. They stated that the mechanism
of drag reduction by spanwise-wall oscillation strongly relates
to the spanwise vorticity where the positive spanwise vorticity
reduced the mean velocity gradient of the boundary layer.
Akhavan et al. [100] reported a 40 % flow enhancement
using oscillating walls. They concluded that spanwise oscillations have been shown to reduce drag more efficiently than
streamwise oscillations in the vast majority of cases. However,
Quadrio and Ricco [102] suggested that the oscillating wall creates a Stokes boundary layer just above the viscous sub-layer;
the Stokes boundary layer holds small localized vortices with
spins opposing the oscillation movement. They stated that the
presence of these small vortices altered the stretching of larger
vortices in the streamwise direction and resulted in a reduction
of the mean vortex diameter which reduced the vorticity.
Recently, Skote [103] introduced a new theory regarding the
drag reduction using oscillating walls. They stated that the
inner part of the turbulent boundary layer is responding
directly to the wall oscillation and the velocity profile is therefore governed by the new (actual) friction velocity. Through
their model derivation, they concluded that the slope of the
logarithmic part of the velocity profile could serve as a measure
of how well the outer part of the flow has adjusted to the
imposed wall forcing.

3.1 Riblet Grooves

Figure 2. Channel flow with multiple dimples at both walls,


case B, fine grid simulation, Re = 10 935 [95].

www.ChemBioEngRev.de

Among all passive means, riblets have been shown to be the


most widely investigated drag reduction technique. Drag
reduction by riblets involves the use of longitudinal microgrooves oriented on a surface originally designed for skin friction reduction in a fully turbulent boundary layer. Originally,
in flow control, it was believed that the smoother a surface, the
longer the delay of the laminar-turbulent transition. However,
in the mid-1960s, Liu et al. [104] and Bath [105] considered
using rough surfaces to reduce the skin friction in turbulent
flow near the wall. By modulating the base flow in the spanwise
direction, these practices were observed to delay the transition
to turbulence.
The early work on riblets focused on riblet geometry and
flow conditions to optimize their drag reduction performance
[43]. In the late 1970s and early 1980s, Walsh and collaborators
at NASAs Langley Research Center started to conduct the first

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

These are not the final page numbers! &&

and Thomas [123] and Savill [106] showed reduced burst freexperimental studies on riblets with turbulent drag reduction
quencies of up to 30 %. Tardu et al. [124] also conducted visuusing several types and shapes of riblets. Since then, a wide
alization studies, which showed that the ejection frequency
variety of interesting studies have emerged and have focused
decreased by 1020 %. Furthermore, Tardu [125] observed that
on five areas: the effect of riblet geometry on drag, the effect of
the riblets shortened the time necessary for the spatio-temporal
heat transfer on wall roughness, the effect of riblet grooves on
development of the instability of low speed streaks in drag
non-Newtonian fluid flow, the effect of riblet grooves on lamireducing configurations. Baron and Quadrio [126] confirmed
nar and transition flows, and simulation analyses and numerithat the duration of the ejections tended to be reduced over the
cal modeling studies of riblets in several applications. Moreriblets and that the reduction compensated for the observed
over, many reviews have been conducted, e.g., Savill [106],
increased burst frequency. Furthermore, they observed that the
Wilkinson et al. [42], Vukoslavcevic et al. [107], Nieuwstadt
ejections tended to be more closely grouped over the riblets.
et al. [108], and Walsh et al. [109, 110].
Bacher and Smith [122] investigated the interactions of the
In 1990, Walsh [43] authored a comprehensive review about
counter rotating longitudinal vortices with the resultant small
the research on drag reduction using riblets on flat-plate turbueddies near the peaks of the riblets. They claimed that the seclent boundary-layer flows. According to Walsh [43], the approondary vortices would act to weaken the longitudinal vortices
priate riblet dimensions for drag reduction vary depending on
and maintain the fluid at low speeds within the grooves.
their shape and the state of the boundary layer in which they
Bechert and his co-authors [112, 121, 127, 128] performed
are placed. Ji-sheng and Heng [111] showed in their study that
theoretical and experimental investigations on triangular, trathe coherent structure of the wall region directly influenced the
pezoidal, blade, semi-circular scalloped, and convex riblets in
production of turbulence energy. The authors concluded that
oil and wind channels. The highest drag reduction percentage
manipulating this structure can reduce the drag force.
was 8.7 % with a riblet spacing of between approximately 2 and
The drag reduction mechanisms of fluid flow over a rib sur10 mm in the oil channel. They focused their investigations on
face are fairly good understood because the structure of turbuthe longitudinal and cross-flow directions. They suggested a
lence flow is at presence well investigated due to numerous themechanism dependent on protrusion heights, where the virtual
oretical, experimental and computational works [112114].
origin of a riblet surface is located at the extrapolation of the
However, it is well known that rib surfaces only rearrange the
streamwise velocity profile. The difference between this origin
turbulence structures near the wall and only influence that
and the riblet peak was defined as the protrusion height for the
localized region near the wall. Additionally, longitudinal ribs
longitudinal flow (hpl). Then, by extrapolation of the spanwise
can develop a lower shear stress flow than a smooth surface
velocity profile, they showed that there was a second protrusion
[43, 115, 116].
height for the cross flow (hpc). The difference in size of these
The influence of riblets on the development and structure of
two heights was dependent on the ratio of height h and spacing
a boundary layer flow is an important research area. Extensive
of the riblets s where the maximum possible difference was
work has been performed by Choi [117], Baron and Quadrio
0.132 s. The theory was based on the hypothesis that the riblets
[118], Park and Wallace [119], and Suzuki and Kasagi [120].
impeded the instantaneous cross-flow in the viscous sublayer,
A grooved surface (riblets) impedes the cross flow or the
which was generated by the turbulent motion, and reduced the
direction of the flow closest to the wall. Therefore, a grooved
momentum and the wall shear stress. The net effect of the
surface prevents random low speed streaks from converging
reduced turbulence intensity was a thickening of the viscous
into an ejection. The riblet protrusion height is an important
sublayer, which translates into a reduction of the wall shear
parameter and is considered to be the offset between the virtual
stress. Fig. 3 shows schematics of the mechanism of fluid flow
origin in the streamwise shear flow and some mean surface
over the longitudinal and cross-flow directions on a ribbed surlocation. Compared with a smooth wall, this offset would result
face.
in a greater separation between the wall and the turbulent
Choi [117] suggested that drag reduction occurs due to the
streamwise vortices, reducing the momentum exchange at the
structural changes in the near-wall boundary layer caused by
wall [121].
the riblets. The authors explained that the riblets would funcSavill [44] suggested that riblets have two drag reducing
tion as small fences to inhibit the lateral turbulence near the
mechanisms. First, riblets shift the turbulence away from the
wall surface. This phenomenon would also limit the lateral
wall, and second, riblets buffer the spacing between the structures to reduce the vicinity of the maximum reduction position. Researchers
a)
b)
have conflicted observations on burst
mechanisms, which plays an important
part in turbulence development [70].
Hooshmand et al. [116] observed an increase in burst frequency and found
that the streamwise velocity was independent of the spanwise position far
away from the surface. However, Walsh
[109] and Bacher and Smith [122]
found no changes due to bursts. VisualFigure 3. (a) Longitudinal and (b) cross-flow on a ribbed surface [112].
ization studies conducted by Gallagher

www.ChemBioEngRev.de

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

These are not the final page numbers! &&

movement of the hairpin legs to the riblets during the near-wall


bursts and lead to premature bursts. When these happen, the
intensity and duration of the subsequent bursts decreases. The
spanwise correlation length increases because the bursts occur
prematurely before the legs approach one another. The frequency of the bursts is increased over the riblet surface. Another hypothesis was suggested by Djenidi et al. [129] when low
speed streaks were observed to exist near the wall in turbulent
boundary layers. These streaks elongated in the streamwise
direction before lifting off the surface. This result was hypothesized to be caused by low speed regions inside the valleys of the
riblets. Drag reductions in laminar flow over riblets should also
be expected. Tang and Clark [124], introduced the concept of a
riblets sublayer. Within this layer, the flow characteristics were
significantly reformed as a result of the presence of the riblets;
additional flow features existed which were unique to the layer.
Three types of secondary eddy patterns formed in the sporadic
cross-flow regions across the riblet peaks, in the longitudinal
near-wall vortices that entered the riblet grooves, and in the
sweep flow from the log-layer that further entrained flow into
the riblet grooves.
Choi et al. [130] conducted a numerical study on laminar
and turbulent channel flows. Choi discussed the drag reduction
mechanisms in terms of the increased spanwise effective viscosity and the thickened viscous sublayer. The rib spacing reduced
the viscous drag by restricting the location of the streamwise
vortices above the wetted surface. He suggested a conceptual
near-wall region model which involved the sweep of high momentum fluid toward the wall surface between pairs of merging
counter-rotating longitudinal vortices. Fig. 4 shows schematic
diagrams of drag increasing and reducing mechanisms. Supposedly, the riblets would impede the lateral movement of the
longitudinal vortices during the near-wall sweep, leading to a
shorter and lower pre-mature sweep. However, the mechanism
of how this process would occur was not provided. Furthermore, the proposed mechanism relied on the implication that
longitudinal vortices near the wall surface were essentially
paired, even though experimental evidence in support of this
inference was lacking.

a)

b)

Figure 4. Schematic diagram of drag increasing and reducing


mechanisms of riblets: (a) s+ = 40 drag increase (extensive area
affected by downwash motion); (b) s+ = 20 drag reduction (limited area affected by downwash motion)[130].

Park and Wallace [119] conducted a comprehensive investigation on the spacing between riblets, i.e., V-groove valleys.
They concluded that drag reduction was primarily derived

www.ChemBioEngRev.de

from greatly reduced wall shear stresses near the bottom of the
riblet valley. The authors noted that the turbulence intensities
of all three velocity components were reduced in the riblet valleys. The intensities increased over the peaks when the height
was less than the viscous sublayer thickness.
Wang et al. [131] and Ma et al. [132] found in a visualization
study that riblets effectively reduced the number of the low-speed
streaks and suppressed the formation of vortices (i.e., hairpin
vortices, ring vortices and horseshoe vortices) and thereby reduced the frequency and intensity of turbulence bursts.
Boiko et al. [133] conducted an experimental investigation
on the effect of riblets on streaky flows and found that riblets
have the ability to reduce the velocity gradient of streamwise
streaky flows in the boundary layer.

3.2 First Inspiration in Nature


Nature is full of flow control examples and has long provided
subjects of interest in the field of fluid mechanics. From birds
flying in a V-formation to feathers on a bird, to scales on a butterfly wing, to the skin of marine creatures, such as octopus,
fish, dolphins, and sharks, natural solutions to aerodynamics
and hydrodynamics are contrasted from the hard, smooth surfaces employed in most engineering applications. Fig. 5 shows
surface structure examples from nature. However, relatively little quantitative research has been performed to determine the
benefits that could be gained by the use of these biological surface patterns due to their often complex three-dimensional
nature. With the use of flow measurement techniques, such as
particle image velocimetry (PIV), it has become possible to analyze fluid flow over such surface geometries and compare the
results with those of more conventional surfaces [134137].
As mentioned by Reif [138, 139], the concept of using riblets
to reduce skin friction drag first came from the observation of
the swimming technique of sharks. Scientists have known that
sharks are one of the fastest swimming animals with burst
speeds of over 20 m s1 [140].
Sharks are extremely agile and maneuverable, capable of instantly changing their direction even at high speeds. The body
of a shark has a sleek, torpedo-like profile to minimize drag.
Additionally, it has been suggested that the scales covering the
shark could be a source of further drag reduction and flow separation control. A shark scale contains grooves that run parallel
to the flow direction and act as riblets over its surface, thereby
decreasing drag by deterring cross flow [128].
Fish [141] examined potential mechanisms of drag reduction
used by dolphins. He noted that the dolphin body shape is that
of a highly streamlined body, similar to a submarine design,
and is a likely factor contributing to the lower than expected
drag. However, Fish noted that during observations of dolphins
swimming, little to no separation showed over their bodies,
whereas fluid flow over duplicated dolphin body shapes were
observed to show flow separations on the surfaces. The author
noted that their body shape cannot be solely responsible for
their efficiency. It is theorized that dolphin skin also acts as a
compliant wall, employing viscous damping to reduce drag.
Their skin may absorb perturbations in the boundary layer that
would ordinarily lead to a transition to turbulent flow. By

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

These are not the final page numbers! &&

Figure 5. Surface structure examples from nature.

delaying this transition, the dolphin skin reduces drag over its
surface [141].

Table 1. Riblet models tested by NASA [142].


Model

Experimental Investigation over


Riblets in Channels

It has long been acknowledged that the manipulation of the


turbulence boundary layer by riblet grooves enhances drag
reduction, and that the riblet geometries (i.e., rib spacing s and
rib height h) affect drag reduction.
In this comparative study, the most suitable riblet type is
defined as one that can offer a high percentage of drag reduction. The comparison was based on the percentage of drag
reduction of different riblet types having different groove geometries. The selection of the riblet type best suited for drag
reduction was determined by the coherent structure of the turbulent flow, the physical properties of the fluid and the riblet
geometries and shape.
Many different types of riblets have been used over the past
four decades, each of which has been geometrically optimized
to achieve higher drag reduction percentages. Tab. 1 shows
some of the riblets shapes that have been tested by NASA.
For viscous drag reductions over a flat plate at low speed
flows, Walsh [109] identified the riblet height and spacing in
+
terms of wall units as non-dimensional
rheight h and spacing
r
Cf
h u
s u C f
+

s , which are defined as h


and s
,
n
2
n
2
respectively, where Cf is the local skin friction coefficient, u is
the free-stream velocity in m s1, h is the kinematic viscosity in
m2s1, s is the peak to peak spacing of riblets and h is the valley
to peak height of riblets.
Investigations have revealed the features of surface textures
in many different applications, such as enhancing flow in pipes
and reduce the drag surfaces, self-cleaning types of surface,
deicing, on commercial aeroplanes, fighter planes, underwater

www.ChemBioEngRev.de

Description
Symmetric V-groove

Rectangular

Spaced triangular

Right angle rib

Peak curvature
Valley curvature
Peak and valley curvature

Notched peak
Spaced V-groove
Unsymmetrical groove

Oblique V-groove

vehicles and others applications [143150]. With this wide


range of applications, there are significant worldwide interests
in this field.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

These are not the final page numbers! &&

Nonetheless, several challenges face the development of surface textures and their translation into industrial applications.
No single universal riblets design can be used for all applications due to completely random turbulent regime structures
and fluid to fluid differences of the viscous sublayers and
streamwise vortices. For example, the ideal riblet structure for
gas turbines compressor blades have spacings of 20 to 120 mm
[121, 151153]. Furthermore, the ability to accurately manufacture riblets is another field of study as there are plenty of
techniques that may be used to fabricate micro-grooves
[154, 155]. Other challenges include determining the appropriate material type and optimizing the type of fabrication process, such as rolling, grinding, etching, and laser etching [136].
NASA machined the first riblets in the late 1970s. These riblets were made from aluminum sheets for aeroplane applications. Later in the 1980s, NASA started using vinyl rib films
manufactured by 3M because of its more cost-effective design.
A polyvinylidene fluoride-based thin plastic film with an adhesive backing was used in the manufacture of micro-patterned
riblets by grinding and rolling processes [156158]. 3M vinyl
rib films have been used for many test surfaces and various
applications, such as aerospace surfaces, submerged vehicles,
and pipes and channels [159167]. In addition, one of the
advantages of the 3M vinyl rib film is the flexibility of the thin
plastic sheets film which can be applied in nearly any orientation, except within small pipes or micro channels, which would
need to be split open to apply the 3M vinyl rib films [136].

4.1 Riblets in Air Flows

tion revealed that the low-speed streak in the riblet valleys


seemed to disappear. Additionally, the data showed 2025 %
burst frequency reductions, depending on the spacing of the
ribs in wall variables. There was a potential drag reduction of
up to 3 %; however, the authors did not feel that their data indicated sufficient turbulence suppression to result in significant
overall drag reduction.
In an experimental study by Furuya et al., conducted over
rectangular bars placed on a flat plate [170], a series of turbulence characteristics were measured (i.e., velocity, turbulence
intensity and wall shear stress distributions). The study showed
that the behavior of the flow was not influenced by changes in
the surface roughness except in the region relatively near the
leading edge. The vortex measurements indicated that the
length scale of the secondary flow vortex grew in accordance
with the boundary layer development, while the vorticity,
which is more intense in the edge region, decreased in that
direction. The total frictional drag of the plate on which a
streamwise bar is placed was almost the same as that for a flat
plate having the same surface area exposed to the flow.
Walsh and Weinstein [171] examined the effects of micro
variations in the surface geometry on altering the near-wall
structure of the turbulent boundary layer. Rectangular and triangular-shaped riblets with different geometries were used.
Fig. 6 shows the dimensions of the tested ribbed surfaces.
The results showed that no net drag reductions were observed for the rectangular ribs. However, the triangular ribs
had smaller drag increases. The drag decreased when the spacing between the triangular ribs was reduced, resulting in a
V-groove type configuration. The highest percentage of drag
reduction, 4 %, was recorded for V-groove riblets with 1520
height and spacing values in the law of the wall coordinates of
the ribs. In a later study, Walsh [172] tested additional
V-groove type configurations to optimize the spacing and
heights of the grooves and to determine the effect of a free

Many studies have focused on wind drag force reductions for


aeroplanes by applying particular groove types and sizes on
aeroplane bodies. NASA Langley Research Center, Boeing and
Airbus have tested aircraft with riblets to investigate their influence on reducing the drag force [20, 168, 169].
Because of the high costs in fabricating and testing
riblets on aeroplanes, riblets have been applied on
reasonably-sized flat plates for their study in wind
tunnels. Thus, different riblet geometries and types
could be studied for their effects on turbulent wind
flow. Experimental studies conducted in this manner are listed in Tab. 2.
From the 1960s to mid-1980s, most of the investigations were focused on determining the effect of
riblet shapes and dimension on suppressing nearwall turbulence. As mentioned earlier, Liu [104],
was the first who showed that an increase in surface roughness suppressed the near-wall turbulence
and confined the turbulent regime by using transverse roughness elements made of square bars, with
ranges of h+ = 45110 and s+ = 190370. The most
obvious effect of an increase in the surface roughness was the distinct increase of turbulence production and the concomitant increase in the eddy viscosity. It was shown that this increase in eddy
viscosity increased the total thickness of the layer
in the same manner that an increase in molecular
Figure 6. Rectangular and triangular ribbed surfaces tested by Walsh and Weinviscosity would do in a laminar flow. The investigastein [171].

www.ChemBioEngRev.de

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

These are not the final page numbers! &&

Table 2. List of experimental wind studies over flat plates with different types of riblets.
Original authors

Year

Riblet type

Riblet dimensions

Drag reduction percentage

Liu [104]

1966

Square

h = 45110
s+ = 190370

3%

Furuya et al. [170]

1977

Rectangular

h = 60 000 mm
s = 20 000 mm

No drag reduction

Walsh and Weinstein [171]

1979

V-groove

h+ = 25, s+ = 20
h = 510 mm, s = 250 mm

24 %

Walsh [172]

1983

V-groove

h+ = 10
s+ = 15

8%

Walsh and Lindemann [110]

1984

V-groove

h+ = 13
s+ = 15

78 %

Enyutin et al. [173]

1987

Triangular

h = 170 mm
s = 250 mm

78 %

Rectangular

h = 215 mm
s = 260 mm

79 %

Wilkinson and Lazos [175]

1988

Rectangular

h+ =8
s+ = 10

8.5 %

Choi [176]

1988

Trapezoidal groove

h+ = 12, s+ = 20
h = 1500 mm, s = 2500 mm, t =
200 mm

4%

Nguyen et al. [156]

1990

V-groove

h+ = 8, s+ = 20
h = 200 mm, s = 500 mm

5%

Enyutin et al. [174]

1991

V-groove

h = 500 mm, s = 260 mm


h/s = 0.3

810 %

Baron and Quadrio [118]

1993

V-groove

h+ = s+ = 12
h = s = 700 mm

6%

Nieuwstadt et al. [108]

1993

V-groove

h+ = s+ = 13
h = s = 640

5%

Park and Wallace [119]

1994

V-groove

h+ = 14
s+ = 28

4%

Gudilin et al. [54]

1995

V-groove

s+ = 1215, a = 53
h = 140 mm, s = 320 mm

89 %

Choi and Orchard [177]

1997

V-groove

h = s = 730 1830 mm

6%

Bechert et al. [128]

2000

Trapezoidal groove

6.85 %

s = 16, a = 45

Lee and Lee [7]

2001

Semi-circular

s = 25.2

Drag decrease

Frohnapfel et al. [178]

2007

Trapezoidal groove

h = 150 mm
s = 2h

25 %

Lee and Choi [180]

2008

V-groove

s+ =10.4
(h = 178.6 mm s = 300 mm)

4%

Jovanovic et al. [179]

2010

Trapezoidal groove

h = 150 mm
s = 2h

25 %

stream sweep angle. To improve the riblet drag reduction performance, three rib geometries were tested: a peak curvature, a
valley curvature, and a notched peak V-groove. The results
showed that the maximum drag reduction was 8 % for a
V-groove geometry with a valley curvature and a V-groove
geometry with sharp peaks and valleys with h+ = 812 and

www.ChemBioEngRev.de

s+ = 1520 in the law of the wall coordinates of the ribs. Walsh


and Lindemann [110] verified the law of the wall scaling for
the riblet height and spacing by testing riblets in boundary
layers with larger Reynolds numbers and optimized the riblet
shape to obtain a higher drag reduction percentage. A maximum drag reduction of 78 % was observed for V-groove rib-

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

These are not the final page numbers! &&

lets having h+ = 13 and s+ = 15. The effect of the free stream


sweep angle on drag reduction did not change up to 15, and
no drag reduction was observed at 30. Enyutin et al. [173]
conducted an investigation on two types of riblets (rectangular
and triangular) to evaluate the effect of riblet size and shape on
skin friction drag. The results showed that for rectangular riblets of h = 215 mm and s = 260 mm, the maximum gain was
79 %. For triangular riblets of h = 170 mm and s = 250 mm, the
maximum gain was 78 %. Enyutin and Walsh and Lindemann
used the same type of grooves, but the relative heights of the
riblets h/s were 0.54 and 0.45 in the Walsh and Lindemann
study and 0.68 and 0.8 in the Enyutin study. The friction
advantage observed by Enyutin was greater than that for Walsh
and Lindemann as shown in Fig. 7.

Figure 7. Friction drag reduction with (1) triangular riblets


(h = 170 mm and s = 250 mm) (2) smooth plate, (3) Walsh and Lindemann work [173].

Moreover, Enyutin et al. [174] investigated the influence of


the free stream sweep angle on the drag reduction over different ribbing geometries. The results agreed with the data presented by Walsh and Lindemann, in which they observed only
a small difference (12 %) for a 15 sweep angle.
From the mid-1980s to late-1990s, most experiments were
conducted to understand the characteristics of turbulent flow
over rib surfaces because the coherent structure of turbulence
over a riblet surface is inherently different than that over a
smooth surface. Wilkinson and Lazos [175] studied the effect
of streamwise near-wall thin element riblets under turbulent
flow conditions. The net effects of the thin-element riblets on
turbulent, viscous skin friction resulted in up to 8.5 % drag
reduction, which was similar to other riblet geometries in previous studies.
Choi [117, 176] measured wall pressure fluctuations of turbulent boundary layers over smooth and ribbed surfaces and
observed a substantial increase in the burst frequency but a
drop in burst duration of almost 50 %. Additionally, the fluctuation intensity of the wall shear stress was considerably lower
in the riblet valleys than over flat walls and contained numerous periods of very low quiescent fluctuation intensities.
Nguyen et al. [156] study a symmetrical V-shaped riblet at
riblet angles of 60 and 30, and heights and spacings of

www.ChemBioEngRev.de

200 mm and 500 mm, respectively. Net drag reductions were calculated using momentum thicknesses at the beginning and the
end of both the rib and smooth plates under the same flow
conditions. The highest percentage of drag reduction, 5 %, was
obtained for a free stream velocity of 16 m s1.
Baron and Quadrio [118] performed experiments in a low
speed wind tunnel with triangular grooves of 700 mm height and
700 mm spacing, i.e., h+ = 12 and s+ = 12 wall units. By testing the
velocity profile over the surface, the main velocity profile showed
an upward movement when plotted in the law of the wall structure. This is a well-known characteristic of drag reduction systems and is directly related to an increase in the thickness of the
viscous sublayer. A similar behavior was reported by Choi [117],
in which the author stated that the drag reducing effectiveness of
riblets was not due to only a purely viscous effect, but likely
due also to a direct interaction with the turbulence-producing
structures of the turbulent boundary layer.
The effectiveness of riblets on skin friction drag reduction
under the influence of both zero-pressure and adverse pressure gradients was investigated by Nieuwstadt et al. [108].
Their experiments were conducted using two triangular riblets having the same hight h and space s dimensions (640
and 360 mm). It was found that the maximum skin friction
reduction was 5 % at s+ = 13 with a zero pressure gradient.
However, with an adverse pressure gradient, the results
showed a decrease in the total drag, which indicates there
was no significant difference.
To study the characteristics and effect of the flatness factor on
ribbed surfaces, Park and Wallace [119] measured the streamwise
velocity profiles over a V-groove riblet with dimensions h+ = 14
and s+ = 28 using a hot-wire probe. The results showed that the
maximum drag reduction was 4 % compared with that on a
smooth surface. Grooved riblets reduced the vertical flux of a
streamwise flow momentum within the riblet valleys. In addition, there remains a small probability of a negative axial flow velocity component within the valleys. The vertical flux of streamwise momentum was greatly reduced within the riblet, even for a
dimensionless size somewhat above the drag-reducing range.
Gudilin et al. [54] used a surface with isosceles ribs having a
vertex angle 53, a rib height h = 140 mm, and a distance
between vertices s = 320 mm. The maximum drag reduction for
the ribbed surface was 89 %.
Choi and Orchard [177] studied the effect of riblets on heat
transfer and drag reduction on the lower wall of a flat plate in a
low speed wind tunnel. A triangular riblet with dimensions
s/h = 1 and s = 183 mm was tested. The results showed an
increase of 10 % in the heat transfer coefficient relative to that
of a smooth surface. Also, a 6 % increase in drag reduction was
observed, compared with a smooth wall.
Bechert et al. [128] found that a turbulent boundary layer on
a shark skin surface can help to reduce turbulent shear stress.
The authors duplicated shark skin using three dimensional
trapezoidal riblets to cover 64 % of the surface area of a
400 mm 500 mm test plate. In their study, the highest drag
reduction of 6.85 % was observed at s+ = 16 and a fin height of
h = 0.5 s.
Lee and Lee [7] analyzed the flow field of a turbulent boundary layer inside a semi-circular riblet valley using a smokewire technique for flow visualization and a PIV for velocity

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

10

These are not the final page numbers! &&

measurements. Fig. 8 shows flow visualization images of


streamwise vortices over rib surface, where the investigation
was conducted on surfaces with semi-circular grooves of riblets
spaced at s+ = 25.2 and 40.6 to understand the different drag
decreasing and increasing rib configurations. For drag decreasing configurations, the riblet dimension was h+ = 12.6 and
s+ = 25.2. The experiments showed that the streamwise vortices
and the spanwise motion appeared to be confined above the
riblets and between the tips, respectively. For the drag increasing configurations, the riblet dimension was h+ = 20.3 and
s+ = 40.6, and the experiments showed that the grooves of the
spaced riblets were wider than the average width of the streamwise vortices. Unlike the vortices in the drag reducing configurations, the vortex centers were located between the valleys and
their concentrations were reduced.

4.2 Riblets in Liquid Flows


Using different types of V-groove riblets on a flat plate, a number of water channel experiments observed maximum drag
reductions of approximately 10 % as shown in Tab. 3.
Gallagher and Thomas [123] conducted flow visualizations
to study turbulent boundary layer characteristics over streamwise grooves with dimensions h = s = 1650 mm on a flat plate.
They reported a low skin friction flow in the valleys of the
grooves and a 30 % reduction in the measured bursting frequency for the riblet surface. However, the total drag for both
grooved and smooth surfaces were the same. To determine the
changes in the turbulent flow structure precipitated by the riblet surface relative to a conventional flat-plate, Bacher and
Smith [122, 181] conducted experiments using flow visualizations in a water channel, with non-dimensional V-groove riblet
height and spacing h+ = s+ = 15. They stated that the key to the
riblet drag reduction performance was the difference between
spanwise and streamwise velocity components. The spanwise
flow was always separated over the riblet peaks, while the
streamwise flow was always attached. Fig. 9 shows a streamwise
vortex interaction with riblet surface via viscous effects. Finally,
they concluded that the most effective riblet design for
enhanced drag reduction contains spanwise and streamwise
protrusion heights.

Figure 8. Flow visualization images of streamwise vortices over


rib surface, with (a) drag decreasing configuration and (b) drag
increasing configuration [7].

Frohnapfel et al. [178] and Jovanovic et al. [179] based their


studies on the numerical simulations conducted by Frohnapfel
[178], wherein the drag force reduction results only for imperceptible grooves that were not capable of producing secondary
flows. Several variables determined the groove dimensions,
such as the wall shear velocity and the kinematic viscosity of
the flow medium. Based on this study, the square cross-section
gave a maximum drag reduction of 33 %. The spacing between
the grooves was designed to avoid local peaks of high wall
shear stresses close to the edges of the grooves. The minimum
groove dimension suggested was h 150 mm, with a separation
of 2 h 300 mm. The suggested grooves were tested; the maximum drag reduction observed was up to 25 % compared with
that of flow over a smooth wall at low Reynolds numbers.
Lee and Choi [180] investigated the effect of spanwise vortices in a turbulent flow over V-groove riblets using a PIV system. The authors visualized the sectional images of the streamwise vortices and measured the velocity fluctuation over the
riblet surface. The riblet had dimensions of 176.8 mm height
and 300 mm spacing between each V-groove riblet.
The results show that small-scale spanwise vortices were
reduced in the near-wall region, and therefore, the thickness of
the viscous sublayer increased.

www.ChemBioEngRev.de

Figure 9. Schematic of streamwise vortex interaction with riblet


surface via viscous effects [181].

Reidy [160] and Reidy and Anderson [159] evaluated the


drag reduction properties of riblets in turbulent boundary
layers. The first experiment used V-groove riblets over a flat
plate in a water tunnel, and the second experiment was conducted in a 6-inch diameter pipeline. 3M vinyl riblets were
used in both systems with a height and spacing of h = s = 76.2
mm. The effects of the riblets on the water turbulent boundary
layer correlated with the previously reported aerodynamic
results, and resulted in a maximum drag reduction of 8.1 %.
Pulles et al. [182] conducted experiments in both water
channels and wind tunnels to study the effect of riblets on the
structure of the turbulent boundary layer. The results showed
that the Reynolds shear stress was noticeably reduced in the
log-law region for a riblet-mounted surface, and the maximum
reduction was approximately 10.6 % for longitudinally-oriented
groove riblets with h = 2500 mm and s = 5000 mm.
Walsh [183] conducted an experiment in a water towing
tank to study the effect of riblet films on turbulence flow, and
the effect of peak and valley curvatures on grooved riblets per-

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

11

These are not the final page numbers! &&

Table 3. List of experimental studies conducted in water on flat plates with different types of riblets.
Original authors

Year

Riblet type

Riblet dimensions
+

Drag reduction percentage

Gallagher and Thomas [123]

1984

V-groove

h = 15
h = s = 1650 mm

No drag reduction

Bacher and Smith [122]

1986

V-groove

h+ = s+ = 15

Drag decrease

Reidy [160]
Reidy and Anderson [186]

1987
1988

V-groove

s = 13.1
h = s = 76.2 mm

8.1 %

Djenidi et al. [129]

1989

V-groove

h+ = s+ = 15
s = h = 2000 mm

4%

Pulles et al. [182]

1989

Triangular

h = 2500 mm

10.6 %

Walsh [183]

1990

V-groove

h = 35.56 mm
s = 40.89 mm

6%

Liu [187]

1990

V-groove

s+ = 13

7%

Neumann and Dinkelacker [184]

1991

Triangular

h = s = 1015
h = s = 152 mm

13 %

Rohr et al. [164]

1992

V-groove

h+ = s+ = 12.2
h = s = 72.6 mm

9%

Parker and Sayers [185]

1999

V-groove

h/s = 0.221

6.83 %

Wang et al. [131]

2000

Trapezoidal grooves

h = 1000 mm
s = 2000 mm

Dean and Bhushan [188]

2012

V-groove

h = 254 mm
h/s = 0.3,0.5,0.7

formance. The V-groove riblets with dimensions h = 35.56 mm


and s = 40.89 mm showed an approximate 6 % drag reduction.
Small deviations in the riblet peak geometry reduced the riblet
drag reduction performance by as much as 40 %, whereas the
valley curvature was not critical to the riblet performance.
Neumann and Dinkelacker [184] performed an experiment
in water with flow velocities of up to 9 m s1 to investigate the
friction drag on a cylindrical body. Triangular-shaped riblets
with h = s = 152 mm dimensions were used. A 9 % turbulent
drag reduction was achieved relative to the flow over a smooth
surface. Additionally, a 13 % drag reduction was observed in
the laminar to turbulent flow transition region. The flow transition on a cylindrical body was delayed when the body was
covered with riblets of a certain size.
Rohr et al. [164] and Liu [169] investigated the performance of 3M vinyl film riblets in internal and external turbulent flows. The maximum drag reduction observed was 9 %
for h+ = s+ = 12.2. No significant change in drag was observed
for the laminar flow region; additionally, the transition was
substantially delayed by riblets.
Parker and Sayers [185] used V-groove riblets of h/s of 0.22
and 1 for h of 0.25 and 0.5 mm, respectively, to reduce the viscous drag on a body. The riblets were machined longitudinally
onto the surface of a smooth plate. The resulting effect on the
drag force of the plate showed that the V-groove riblets reduced
the turbulent skin friction drag by up to 7 %, depending on the
size of the riblets. A boundary layer analysis of the turbulent
flow characteristics over both the smooth and riblet-machined
surfaces indicated an increase in the laminar sublayer thickness

www.ChemBioEngRev.de

and local Reynolds number and a reduced boundary layer


thickness for the ribbed surfaces. A maximum drag reduction
of 6.83 % was recorded for the surface covered with the symmetric riblets (i.e., h/s = 1), at a Reynolds number of 1.17 105
as in Fig. 10.

Figure 10. Direct drag data for the V-groove ribbed, with riblet
surface 1 (h/s = 0.22, h = 0.25 mm) and riblets surface 2 (h/s = 1,
h = 0.5 mm) [185].

The authors concluded that the riblets interrupted the


momentum and turbulent energy exchanges from regions of
high velocity to lower-velocity regions. The riblets impeded the
cross-flow of streamwise vortices that prevailed in the viscous
sublayer of the turbulent boundary layer. By suppressing these

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

12

These are not the final page numbers! &&

streamwise vortices, turbulent mixing and hence, turbulent


shear stresses were reduced.
Several challenges confronted the investigations of drag
reduction over ribbed surfaces in wind and water channels. In
wind channels, because the viscous sublayer was less than
0.1 mm thick, the turbulent shear stress was extremely small
and was difficult to accurately measure by force balance. Additionally, convection issues inside a wind tunnel resulted in poor
experimental conditions. In water channels, the experimental
conditions were similar or worse. At the required very low
speeds, the measuring times became so large that accurate
shear stress measurements were virtually impossible [189].
Because the dimensions of drag reducing structures are as
small as the thickness of the viscous sublayer, a channel with a
liquid viscosity greater than those of water and wind would be
unrealistic to model. Therefore, an oil channel was designed
and fabricated by Bechert et al. [189] to investigate the drag
reduction and the dynamics of the oil viscous sublayer near the
wall on smooth and ribbed surfaces. It was found that the
thickness of the viscous sublayer could be varied between
14 mm. The shear stress data from the oil channel had an
accuracy of +0.2 %. Additionally, the spacing between riblets
could be as large as 310 mm compared with the 0.5 mm rib
spacing for wind channels.
Bechert et al. [112], conducted extensive investigations using
an oil channel to study the effects of the sizes and shapes of riblets and the ratios of the riblet spacing to height (Tab. 4). Baby
oil was used in a rectangular channel (25 cm width and 85 cm
height). The results show that for V-groove riblets with dimensions a = 60, s = 3.034 mm, and s+ = 15, the maximum drag
reduction was approximately 5 %. For the semicircular scalloped riblets, the highest drag reduction was 7.5 % at s+ = 14.
For blade riblets, which are thin plates aligned in the flow direction and perpendicular to the wall, the maximum drag reduction achieved was nearly 10 % at h = 0.5 s.
Bechert et al. [128] similarly tested the channel three-dimensional trapezoidal riblets which covered 64 % of a test plate.
The maximum drag reduction reported was 6.85 % for riblets
with dimensions h/s = 0.85, a = 45, s = 0.46 cm, and s+ = 16.
Two-dimensional riblets produced larger drag reductions than
three-dimensional riblets.
Gruneberger and Hage [190] optimized the shape of riblets
based on results available from many laboratories. Different
shapes, including the triangular, rectangular, trapezoidal, sawtooth, and scalloped cross-sections, were investigated through
experiments and simulations.

Combination Studies of Riblets and


Polymer/Surfactant

In this section, the literature that studied the effects of combining riblets with active means on reducing the drag force is summarized. Reidy and Anderson[159], as mentioned earlier, conducted investigations over a flat plate system and in a pipeline
system with smooth and ribbed surfaces. They also studied the
effect of adding 2 wppm of polyacrylamide to the water flow
over the ribbed surface in the pipeline system. The resultant
maximum drag reduction was 28 %, which is approximately
equal to the sum of the drag reductions of the two techniques
used separately. Rohr et al. [192] conducted experimental studies at low Reynolds numbers with 2.5, 10 and 40 wppm solutions of polyethyleneoxide on smooth and ribbed surfaces in a
pipe. 3M V-groove riblets with dimensions h = s = 76 mm were
used in a 12.7 mm pipe. The riblets had no observable effect on
the polymer solutions, as the calculated drag for the combined
system was the same as the drag reduction ratio of the polymer
solution alone.
Christodoulou et al. [193] investigated the effects of the
combined systems of V-groove riblets with dimensions of
h = s = 110 mm and 250 wppm of Polyox 301 or polyacrylamide in a 25.4 mm pipe. At low concentrations of either polymer, i.e., 210 wppm, and with h+ < 10, the maximum drag
reduction achieved was 4 %. Anderson et al. [165] conducted
an experiment at low polymer concentrations over ribbed
surfaces in turbulent pipe flow. Polyox 301 (at 2, 4, and
8 wppm) and guar gum (at 100 wppm) solutions with
V-groove riblets heights (4 < h+ < 90) were used in this study.
It was found that at low concentrations of polymers, 57 %
drag reduction was observed for a riblet dimension of h+ = 12.
Bewersdorff and Thiel [194] investigated the friction behavior of diluted polymer and surfactant-based DRAs in a smooth
pipe and two rough pipes, with k and d-type roughness. The
surfactant solution exhibited a drag reduction in the smooth
and rough pipes; the polymer and surfactant solutions had no
influence on the roughness of the solutions.
Koury and Virk [195] used aqueous solutions of a polyethyleneoxide polymer (3 to 100 wppm) as a drag reducing agent
with 150 mm wide V-groove riblets in a pipe at Reynolds numbers of 300 to 150 000. It was found that the drag reduction in
the ribbed pipe was 1.6-factors higher than that of smooth
pipes when using 3 wppm polyethyleneoxide polymer. They
observed that with an increasing polymer concentration, the
drag reduction by the riblets decreased to 0.8 for smooth pipes

Table 4. List of experimental studies conducted in oil on flat plates with different types of riblets.
Original authors

Year

Riblet type

Riblet dimensions

Drag reduction percentage

Bechert et al. [112]

1997

Blade riblets

h = 0.5 s

9.9 %

Bechert et al. [128]

2000

Trapezoidal 3D-riblets

s = 16
h/s = 0.85

7.3 %

Gruneberger and Hage [190]

2011

Trapezoidal groove

h+ = 8.5
s+ = 17

7.6 %

Buttner and Schulz [191]

2011

Blade riblets

s+ = 20.5

4.9 %

www.ChemBioEngRev.de

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

13

These are not the final page numbers! &&

(at 10 wppm polymer). The authors concluded that the riblets


and polymers reduced drag by separate mechanisms. In terms
of the turbulent burst cycle, the riblets inhibited the growth
stage and reduced the burst strength, whereas the polymers
altered the breakdown stage and reduced the axial to transverse
energy transfer.
Recent studies were conducted by Chen et al. [196], Zhao
et al. [197], and Zhang et al. [198, 199], in which they proposed
and investigated a new technique based on replication to produce micro-structured, precise, shark skin-like surfaces. The
process prepared replicated fresh shark skin. The best piece of
fresh shark skin was selected based on intact micro-riblets.
Then, the skin was cleaned with distilled water and baked in an
oven at 60 C for 3 h to increase its rigidity. A mold of the shark
skin was then cast using unsaturated polyester resin under vacuum. Multilayer fiberglass was added to prevent distortions in
the riblets and to prevent the mold from cracking. Those investigations showed a 24.6 % maximum drag reduction. Additionally, the group investigated a drag reduction combination of
bioreplicated shark skin and a polymer additive and found that
drag reduction rates reached up to 80 %.

Conclusion

The first inspiration from nature motivated large numbers of


researchers to simulate the effect of the evenly structured shark
skin profile to enhance the flow in conduits, pipes and submerged surfaces as well. Despite all these efforts, the achieved
drag reduction using riblets is still low (around 10 % as an average) and that might be due to the shapes and dimensions optimizations that limit exploring the real effect of such phenomena. One of the difficulties facing the researchers working with
riblets is the cost of conducting experiments covering all the interactive and complex variables involved. This is why a large
number of simulation works (CFD works) where observed in
the literatures using two dimensional models to evaluate the
flow behavior over riblets. Such efforts were successful in introducing close pictures of the real fluid flow behavior over the
structures surfaces, but fail to optimize completely the dimensions-flow relation (which was obvious from the high differences between the experimental and theoretical results). For all
that, the need for the three dimensional modeling is a need at
this stage to narrow the gap between the experimental and theoretical works. The three dimensional model can give a new
insight towards highlighting the effect of the reblits structure
(shape and design) and length which will create a three dimensional turbulence structure and that can visualize clearly the
real effective factors controlling the flow behavior over such
structured surfaces.
It is a fact that most of the instability occurs in the buffer
layer over the structured surfaces and the laminar sub layer
was not examined properly and most of the works done on
these two layers was through the simulation using CFD and
some experimental tests that never revealed the real effect.
This is why it is important to go down to the micro-scale experiments by implementing the micro-channels experiments
where riblets can be built-up using photonic lithography and
the flow observation can be done through microscope PIV.

www.ChemBioEngRev.de

Such an approach will help in understanding the real flow


behavior over the structured surfaces and to suggest real optimized dimensions for the riblets designs especially when
testing liquids with different physical properties (especially
viscosity). This will contribute significantly to the optimization efforts because it will reveal the effect of these properties
on the flow behavior in the micro-channels by eliminating
the flow core bulk effect.

Acknowledgment
This work was supported financially by the University Malaysia
Pahang (UMP) through the Fundamental Research Grant.

Hayder A. Abdulbari is the


director of the Center of Excellence for Advanced Research
in Fluid Flow at the University
Malaysia Pahang. He was appointed to the University of
Baghdad and University of
Technology from 2000 to 2007
and then joined University
Malaysia Pahang in 2007 until
now. He has authored and
published a large number of
papers in the fields of drag
reduction, pipeline flow systems, lubricants, water treatment systems, and emulsions.
The research works carried out by him and his co-researchers was internationally recognized (22 patents) and he
received many international awards from USA, Korea,
Malaysia, Germany and more.
Zulkefli B. Y. Hassan graduated from Camborne School of
Mines with a B.Sc. (Hons) in
Mining Engineering in 1984.
He obtained his M.Sc. in Natural Gas Engineering from
Texas A&M University in
1989. In 1997, he gained his
Ph.D. from Salford University
for his research thesis titled
Modeling and Simulation of
Transient Gas Flow. At the
end of 2011, he started his
tenure at the University Malaysia Pahang in the Faculty of Chemical and Natural
Resources Engineering. He is is currently the Director of the
Industry Partnership and Community Relation of the University and has authored and published numerous papers in
journals and conferences. His research interest include oil
and gas flow technologies.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

14

These are not the final page numbers! &&

Hassan D. Mahammad studied Chemical Engineering at


the University of Technology
Iraq from 20052009 and received his M.Sc. in Bioprocess
Engineering from the University Malaysia Perlis in 2013.
He then moved to Pahang and
started his Ph.D. at the University Malaysia Pahang in the research group of Dr. Hayder
A. Abdulabari in Chemical Engineering. His current research
interest is mostly focused on
drag reduction in pipeline.

Symbols used
h
h+
hpc
hpl
s
s+

[mm]
[]
[mm]
[mm]
[mm]
[]

riblet height
non-dimensional riblet height
height for the cross flow
height for the longitudinal flow
riblet spacing
non-dimensional riblet spacing

Abbreviations
BLAD
DRA
LEBU
OLD
PIV
TAPPM

boundary layer device


drag reducing agent
large eddy breakup device
outer layer device
particle image velocimeter
tandem arrayed parallel plate manipulator

References
[1]
[2]
[3]

[4]
[5]
[6]

[7]
[8]

M. Gad-el-Hak, in Frontiers in Experimental Fluid Mechanics


(Ed: M. Gad-el-Hak), Springer, Berlin 1989, 211290.
J. D. Anderson Jr., Phys. Today 2005, 58 (12), 4248.
M. Gad-el-Hak, Flow Control: Passive, Active, and Reactive
Flow Management, Cambridge University Press, New York
2007.
G. E. Mase, Schaums Outline of Theory and Problems of
Continuum Mechanics, McGraw-Hill, New York 1970.
J. L. Zakin, W. Ge, in Polymer Physics, John Wiley & Sons,
New York 2010, 89127.
J. Cousteix, in Special Course on Skin Friction Drag Reduction (AGARD-R-786) (Ed: J. Cousteix), Advisory Group for
Aerospace Research and Development (AGARD), Neuilly
sur Seine 1992, 139.
S. J. Lee, S. H. Lee, Exp. Fluids 2001, 30 (2), 153166. DOI:
10.1007/s003480000150
F. Gallego, S. N. Shah, J. Petrol. Sci. Eng. 2009, 65 (34),
147161. DOI: 10.1016/j.petrol.2008.12.013

www.ChemBioEngRev.de

[9] N. H. Abdurahman, Y. M. Rosli, N. H. Azhari, B. A. Hayder,


J. Petrol. Sci. Eng. 2012, 9091, 139144. DOI: 10.1016/
j.petrol.2012.04.025
[10] A. P. Szilas, in Developments in Petroleum Science, Elsevier,
Amsterdam 1986, 279340.
[11] J. K. Fink, Oil Field Chemicals, Gulf Professional Publishing,
Burlington, MA 2003.
[12] J. K. Fink, Petroleum Engineers Guide to Oil Field Chemicals
and Fluids, Gulf Professional Publishing, Waltham, MA
2012.
[13] H. A. Abdulbari, R. M. Yunus, N. Norahzan, IEEE Colloquium on Humanities, Science and Engineering (CHUSER),
Kota Kinabalu, December 2012.
[14] H. Abdulbari, N. Kamarulizam, A. H. Nour, Chem. Ind.
Chem. Eng. Quart. 2012, 18 (3), 361371. DOI: 10.2298/
ciceq111206012a
[15] T. Theodorsen, Proc. of the 2nd Midwestern Conf. on Fluid
Mechanics, Ohio State University, OH 1952, 19.
[16] H. E. Fiedler, in Advances in Turbulence (Eds: C. B. Genevie`ve, M. Jean), Springer, Heidelberg 1987, 320336.
[17] R. Falco, in 21st Aerospace Sciences Meeting, American
Institute of Aeronautics and Astronautics, Washington, DC
1983. DOI: 10.2514/6.1983-377
[18] B. A. Toms, in Proc. of the Int. Congr. on Rheology, North
Holland Publisher, Amsterdam 1948, (2), 135141.
[19] M. Gad-el-Hak, Appl. Mech. Rev. 1996, 49 (7), 365380.
DOI: 10.1115/1.3101931
[20] E. Coustols, A. M. Savill, in Special Course on Skin Friction
Drag Reduction (AGARD-R-786) (Ed: J. Cousteix), Advisory
Group for Aerospace Research and Development (AGARD),
Neuilly sur Seine 1992, 8, 154.
[21] V. R. Arunachalam, R. L. Hummel, J. W. Smith, Can.
J. Chem. Eng. 1972, 50 (3), 337343. DOI: 10.1002/
cjce.5450500305
[22] P. S. Virk, AIChE J. 1975, 21 (4), 625656. DOI: 10.1002/
aic.690210402
[23] Y. Wang, B. Yu, J. L. Zakin, H. Shi, Adv. Mech. Eng. 2011, 31,
117. DOI: 10.1155/2011/478749
[24] C. B. Lester, Oil Gas J. 1985, 83 (5), 51.
[25] M. S. N. Kazi, G. G. Duffy, X. D. Chen, Chem. Eng. J. 1999,
73 (3), 247253. DOI: 10.1016/s1385-8947(99)00047-9
[26] A. Al-Sarkhi, T. J. Hanratty, Int. J. Multiphase Flow 2001,
27 (7), 11511162. DOI: 10.1016/s0301-9322(00)00071-9
[27] V. T. Truong, Drag Reduction Technologies, Aeronautical and
Maritime Research Laboratory, Canberra 2001.
[28] C. M. White, M. G. Mungal, Ann. Rev. Fluid Mech. 2008, 40
(1), 235256. DOI: 10.1146/annurev.fluid.40.111406.102156
[29] J. W. Hoyt, in Viscous Drag Reduction in Boundary Layers
(Ed: D. Bushnell), American Institute of Aeronautics and
Astronautics, Washington, DC 1990, 413432.
[30] H. A. Abdulbari, R. M. Yunus, N. H. Abdurahman,
A. Charles, J. Ind. Eng. Chem. 2013, 19 (1), 2736. DOI:
10.1016/j.jiec.2012.07.023
[31] P. Diamond, J. Harvey, J. Katz, D. Nelson, P. Steinhardt, Drag
Reduction By Polymer Additives, Report 22102 JSR-89-720,
US Department of Defense, Washington, DC 1992.
[32] N. P. Cheremisinoff, P. N. Cheremisinoff, Handbook of
Applied Polymer Processing Technology, CRC Press, Boca
Raton, FL 1996.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

15

These are not the final page numbers! &&

[33] S.-H. Cho, C.-S. Tae, M. Zaheeruddin, Energy Convers.


Manage. 2007, 48 (3), 913918. DOI: 10.1016/j.enconman.
2006.08.021
[34] G. A. Agoston, W. H. Harte, H. C. Hottel, W. A. Klemm,
K. Mysels, H. Pomeroy, J. Thompson, Ind. Eng. Chem. 1954,
46 (5), 10171019. DOI: 10.1021/ie50533a055
[35] R. Sharma, Surfactant Adsorption and Surface Solubilization,
American Chemical Society, Washington, DC 1994.
[36] P. C. Hiemenz, R. Rajagopalan, Principles of Colloid and
Surface Chemistry, 3rd revised and expanded ed., CRC Press,
New York 1997.
[37] J. L. Zakin, B. Lu, H.-W. Bewersdorff, Rev. Chem. Eng. 1998,
14 (45), 253320. DOI: 10.1515/revce.1998.14.4-5.253
[38] N. Yusuf, T. Al-Wahaibi, Y. Al-Wahaibi, A. Al-Ajmi, A. R.
Al-Hashmi, A. S. Olawale, I. A. Mohammed, Int. J. Heat
Fluid Flow 2012, 37 (0), 7480. DOI: 10.1016/
j.ijheatfluidflow.2012.04.014
[39] S. J. Rossetti, R. Pfeffer, AIChE J. 1972, 18 (1), 3139. DOI:
10.1002/aic.690180107
[40] R. S. Kane, in Viscous Drag Reduction in Boundary Layers
(Eds: D. M. Bushnell, J. N. Hefner), American Institute of
Aeronautics and Astronautics, Washington, DC 1990,
433456.
[41] R. C. Vaseleski, A. B. Metzner, AIChE J. 1974, 20 (2),
301306. DOI: 10.1002/aic.690200214
[42] S. P. Wilkinson, J. B. Anders, B. S. Lazos, D. M. Bushnell, Int.
J. Heat Fluid Flow 1988, 9 (3), 266277. DOI: 10.1016/0142727X(88)90037-9
[43] M. J. Walsh, in Viscous Drag Reduction in Boundary Layers
(Eds: D. M. Bushnell, J. Hefner), American Institute of
Aeronautics and Astronautics, Washington, DC 1990,
203261.
[44] A. M. Savill, in Structure of Turbulence and Drag Reduction,
Springer, Heidelberg 1990, 429465.
[45] M. R. Raupach, R. A. Antonia, S. Rajagopalan, Appl. Mech.
Rev. 1991, 44 (1), 1. DOI: 10.1115/1.3119492
[46] G. E. A. Meier, G. H. Schnerr, E. Coustols, in Control of Flow
Instabilities and Unsteady Flows, Springer, Berlin 1996,
155202.
[47] D. M. Bushnell, Proc. Inst. Mech. Eng., Part G 2003, 217 (1),
118. DOI: 10.1243/095441003763031789
[48] F. E. Fish, G. V. Lauder, Ann. Rev. Fluid Mech. 2006, 38 (1),
193224. DOI: 10.1146/annurev.fluid.38.050304.092201
[49] A. R. Paul, S. Joshi, A. Jindal, S. P. Maurya, A. Jain, Sci.
World J. 2013, 2013. DOI: 10.1155/2013/523759
[50] E. Coustols, A. M. Savill, in Special Course on Skin Friction
Drag Reduction (Ed: J. Cousteix), Advisory Group for Aerospace Research and Development (AGARD), Neuilly sur
Seine 1992, 287.
[51] K. Watanabe, T. Takayama, S. Ogata, S. Isozaki, AIChE J.
2003, 49 (8), 19561963. DOI: 10.1002/aic.690490805
[52] A. Hamdouni, J. P. Bonnet, Appl. Sci. Res. 1993, 19 (34),
369385. DOI: 10.1007/978-94-011-1701-2_11
[53] A. Pollard, Prog. Aerosp. Sci. 1998, 33 (1112), 689708.
DOI: 10.1016/S0376-0421(97)00008-0
[54] I. V. Gudilin, Y. A. Lashkov, V. G. Shumilkin, Fluid Dyn.
1995, 30 (3), 366371. DOI: 10.1007/bf02282448

www.ChemBioEngRev.de

[55] J. Anders, J. Hefner, D. Bushnell, in 22nd Aerospace Sciences


Meeting, American Institute of Aeronautics and Astronautics, Washington, DC 1984.
[56] M. J. Walsh, J. B. Anders, Appl. Sci. Res. 1989, 46 (3),
255262. DOI: 10.1007/bf00404822
[57] A. Sahlin, A. V. Johansson, P. H. Alfredsson, Phys. Fluids
1988, 31 (10), 28142820. DOI: 10.1063/1.866989
[58] T. B. Lynn, D. A. Gerich, D. W. Bechert, in Advances in
Turbulence 3 (Eds: A. Johansson, P. H. Alfredsson), Springer,
Berlin 1991, 472480.
[59] P. R. Spalart, M. Strelets, A. Travin, Int. J. Heat Fluid Flow
2006, 27 (5), 902910. DOI: 10.1016/j.ijheatfluidflow.
2006.03.014
[60] B. Eckhardt, V. I. Kornilov, A. V. Boiko, in Advances in
Turbulence XII (Ed: B. Eckhardt), Springer, Berlin 2009,
205208.
[61] B. Vasudevan, A. Prabhu, R. Narasimha, Exp. Fluids 1992,
12 (3), 200208. DOI: 10.1007/bf00188259
[62] J. Gray, J. Exp. Biol. 1936, 13 (2), 192199.
[63] M. O. Kramer, J. Am. Soc. Nav. Eng. 1960, 72 (1), 2534.
DOI: 10.1111/j.1559-3584.1960.tb02356.x
[64] M. Gad-el-Hak, Exp. Therm. Fluid Sci. 1998, 16 (12),
141156. DOI: 10.1016/S0894-1777(97)10006-1
[65] S. Xu, D. Rempfer, J. Lumley, J. Fluid Mech. 2003, 478 (1),
1134. DOI: 10.1017/s0022112002003324
[66] F. W. Puryear, Boundary Layer Control-drag Reduction by
Use of Compliant Coatings, David Taylor Model Basin Report, National Aeronautics and Space Administration,
Washington, DC 1962.
[67] C. R. Nisewanger, Flow Noise and Drag Measurements of Vehicle with Compliant Coating, U.S. Naval Ordenance Test
Station China Lake, CA 1964.
[68] H. Ritter, L. T. Messum, Water Tunnel Measurements of Turbulent Skin Friction on Six Different Compliant Surfaces of
1 Ft Length, ARL/N4/G/HY/9/7, British Admiralty Research
Laboratory, London 1964.
[69] A. O. Steven, Prediction of Compliant Wall Drag Reduction,
Part II, Technical Report, National Aeronautics and Space
Administration (NASA), Washington, DC 1979.
[70] S. J. Kline, W. C. Reynolds, F. A. Schraub, P. W. Runstadler,
J. Fluid Mech. 1967, 30 (4), 741773. DOI: 10.1017/
s0022112067001740
[71] M. Fischer, L. Weinstein, D. Bushnell, R. Ash, in 8th Fluid
and PlasmaDynamics Conference, American Institute of
Aeronautics and Astronautics, Washington, DC 1975.
[72] P. W. Carpenter, A. D. Garrad, J. Fluid Mech. 1985, 155, 465.
DOI: 10.1017/s0022112085001902
[73] P. W. Carpenter, A. D. Garrad, J. Fluid Mech. 1986, 170,
199232. DOI: 10.1017/S002211208600085X
[74] M. Gad-el-Hak, Prog. Aerosp. Sci. 2002, 38 (1), 7799. DOI:
10.1016/s0376-0421(01)00020-3
[75] M. L. McMillan, H. C. Hershey, R. A. Baxter, in Chemical
Engineering Progress Symposium Series, American Institute
of Chemical Engineers, New York 1971, 2744.
[76] M. Rodrguez, J. Xue, L. M. Gouveia, A. J. Muller, A. E. Saez,
J. Rigolini, B. Grassl, Colloids Surf. A 2011, 373 (13), 6673.
DOI: 10.1016/j.colsurfa.2010.10.024
[77] E. Soali, A. B. Hayder, Z. Hasan, M. Rahman, J. Appl. Sci.
2010, 10 (21), 26832687. DOI: 10.3923/jas.2010.2683.2687

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

16

These are not the final page numbers! &&

[78] A. Krope, L. C. Lipus, Appl. Therm. Eng. 2010, 30 (89),


833838. DOI: 10.1016/j.applthermaleng.2009.12.012
[79] S. Hofmann, P. Stern, J. Myska, Rheol. Acta 1994, 33 (5),
419430. DOI: 10.1007/bf00366584
[80] I. Harwigsson, M. Hellsten, US Patent WO 1996028527 A1,
1996.
[81] H. A. Abdulbari, R. B. M. Yunus, T. S. Hadi, Am. J. Appl. Sci.
2010, 7 (10), 13101316. DOI: 10.3844/ajassp.2010.1310.
1316
[82] Y. Zhang, J. Schmidt, Y. Talmon, J. L. Zakin, J. Colloid Interface Sci. 2005, 286 (2), 696709. DOI: 10.1016/j.jcis.2005.
01.055
[83] J. J. Wei, Y. Kawaguchi, F. C. Li, B. Yu, J. L. Zakin, D. J. Hart,
Y. Zhang, Int. J. Heat Mass Transfer 2009, 52 (1516),
35473554. DOI: 10.1016/j.ijheatmasstransfer.2009.03.008
[84] R. Pal, AIChE J. 1993, 39 (11), 17541764. DOI: 10.1002/
aic.690391103
[85] H. K. Moon, T. OConnell, B. Glezer, J. Eng. Gas Turbines
Power 2000, 122 (2), 307313. DOI: 10.1115/1.483208
[86] P. M. Ligrani, J. L. Harrison, G. I. Mahmmod, M. L. Hill,
Phys. Fluids 2001, 13 (11), 3442. DOI: 10.1063/1.1404139
[87] G. I. Mahmood, P. M. Ligrani, Int. J. Heat Mass Transfer
2002, 45 (10), 20112020. DOI: 10.1016/S0017-9310(01)
00314-3
[88] S. Y. Won, Q. Zhang, P. M. Ligrani, Phys. Fluids 2005, 17 (4),
DOI: 10.1063/1.1872073
[89] P. M. Ligrani, G. I. Mahmood, J. L. Harrison, C. M. Clayton,
D. L. Nelson, Int. J. Heat Mass Transfer 2001, 44 (23),
44134425. DOI: 10.1016/S0017-9310(01)00101-6
[90] M. A. Elyyan, A. Rozati, D. K. Tafti, Int. J. Heat Mass Transfer 2008, 51 (1112), 29502966. DOI: 10.1016/j.ijheatmasstransfer.2007.09.013
[91] T. S. Griffith, L. Al-Hadhrami, J.-C. Han, J. Turbomach.
2003, 125 (3), 555563. DOI: 10.1115/1.1571850
[92] W. Rohlfs, H. D. Haustein, O. Garbrecht, R. Kneer, Int. J.
Heat Mass Transfer 2012, 55 (2526), 77287736. DOI:
10.1016/j.ijheatmasstransfer.2012.07.081
[93] B. Yu, F. Li, Y. Kawaguchi, Int. J. Heat Fluid Flow 2004, 25
(6), 961974. DOI: 10.1016/j.ijheatfluidflow.2004.02.029
[94] Z. Xu, S. Z. Li, X. Y. Wu, X. J. Zhao, Adv. Mater. Res. 2011,
299300, 711. DOI: 10.4028/www.scientific.net/AMR.299300.7
[95] H. Lienhart, M. Breuer, C. Koksoy, Int. J. Heat Fluid Flow
2008, 29 (3), 783791. DOI: 10.1016/j.ijheatfluidflow.2008.
02.001
[96] Y. Du, V. Symeonidis, G. E. Karniadakis, J. Fluid Mech. 2002,
457, 134. DOI: 10.1017/s0022112001007613
[97] A. Baron, M. Quadrio, Appl. Sci. Res. 1996, 55 (4), 311326.
DOI: 10.1007/bf00856638
[98] M. Quadrio, P. Ricco, J. Fluid Mech. 2014. 521, 251271.
DOI: 10.1017/S0022112004001855
[99] N. V. Nikitin, Fluid Dyn. 2000, 35 (2), 185190. DOI:
10.1007/BF02831425
[100] F. T. M. Nieuwstadt, R. Akhavan, W. J. Jung, N. Mangiavacchi, in Advances in Turbulence IV, Springer, Berlin 1993,
299303.
[101] K.-S. Choi, B. R. Clayton, Int. J. Heat Fluid Flow 2001, 22
(1), 19. DOI: 10.1016/s0142-727x(00)00070-9

www.ChemBioEngRev.de

[102] M. Quadrio, P. Ricco, J. Fluid Mech. 2004, 521, 251271.


DOI: 10.1017/s0022112004001855
[103] M. Skote, Int. J. Heat Fluid Flow 2014, 50 (0), 352358. DOI:
10.1016/j.ijheatfluidflow.2014.09.006
[104] C. K. Liu, Ph.D. Thesis, Stanford University, CA 1966.
[105] T. D. Bath, Channeled Flow at the Pipe Surface in Gas Transmission Pipelines, Midwest Research Institute, Kansas City,
MO 1968.
[106] A. M. Savill, in Flow Visualization IV: Proc. of the 4th Int.
Symp. on Flow Visualization, Springer, Berlin 1987, 303
308.
[107] P. Vukoslavcevic, J. M. Wallace, J. L. Balint, AIAA J. 1992, 30
(4), 11191122. DOI: 10.2514/3.11035
[108] F. T. M. Nieuwstadt, W. Wolthers, H. Leijdens, K. Krishna
Prasad, A. Schwarz-van Manen, Exp. Fluids 1993, 15 (1),
1726. DOI: 10.1007/BF00195591
[109] M. J. Walsh, in 20th Aerospace Sciences Meeting, American
Institute of Aeronautics and Astronautics, Washington, DC
1982.
[110] M. Walsh, A. Lindemann, in 22nd Aerospace Sciences
Meeting, American Institute of Aeronautics and Astronautics, Washington, DC 1984.
[111] L. Ji-sheng, Z. Heng, Appl. Math. Mech. 1993, 14 (11),
9931001. DOI: 10.1007/bf02476547
[112] D. W. Bechert, M. Bruse, W. Hage, J. G. T. Van Der Hoeven,
G. Hoppe, J. Fluid Mech. 1997, 338, 5987. DOI: 10.1017/
s0022112096004673
[113] P. Luchini, F. Manzo, A. Pozzi, J. Fluid Mech. 1991, 228,
87109. DOI: 10.1017/S0022112091002641
[114] S. K. Robinson, Ph.D. Thesis, Stanford University, CA 1991.
[115] P. Nitschke, Experimental investigation of the turbulent flow
in smooth and longitudinal grooved tubes, NASA TM77480,
National Aeronautics and Space Administration, Washington, DC 1983.
[116] D. Hooshmand, R. Youngs, J. M. Wallace, J. L. Balint, in 21st
Aerospace Sciences Meeting, American Institute of Aeronautics and Astronautics (AIAA), Washington, DC 1983. DOI:
10.2514/6.1983-230
[117] K.-S. Choi, J. Fluid Mech. 1989, 208 (1), 417. DOI: 10.1017/
s0022112089002892
[118] A. Baron, M. Quadrio, Int. J. Heat Fluid Flow 1993, 14 (3),
223230. DOI: 10.1016/0142-727X(93)90052-O
[119] S.-R. Park, J. M. Wallace, AIAA J. 1994, 32 (1), 3138. DOI:
10.2514/3.11947
[120] Y. Suzuki, N. Kasagi, AIAA J. 1994, 32 (9), 17811790. DOI:
10.2514/3.12174
[121] D. W. Bechert, M. Bartenwerfer, J. Fluid Mech. 1989, 206 (1),
105129. DOI: 10.1017/S0022112089002247
[122] E. V. Bacher, C. R. Smith, AIAA J. 1986, 24 (8), 13821385.
DOI: 10.2514/3.48695
[123] J. Gallagher, A. Thomas, in 2nd Applied Aerodynamics
Conference, American Institute of Aeronautics and Astronautics, Washington, DC 1984, 9.
[124] S. Tardu, T. V. Truong, B. Tanguay, Appl. Sci. Res. 1993, 50
(34), 189213. DOI: 10.1007/BF00850557
[125] S. F. Tardu, Appl. Sci. Res. 1995, 54 (4), 349385. DOI:
10.1007/bf00863518
[126] A. Baron, M. Quadrio, Int. J. Heat Fluid Flow 1997, 18 (2),
188196. DOI: 10.1016/S0142-727X(96)00087-2

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

17

These are not the final page numbers! &&

[127] D. W. Bechert, Int. Conf. on Turbulent Drag Reduction by


Passive Means, London, September 1987.
[128] D. W. Bechert, M. Bruse, W. Hage, Exp. Fluids 2000, 28 (5),
403412. DOI: 10.1007/s003480050400
[129] L. Djenidi, J. Liandrat, F. Anselmet, L. Fulachier, in Advances
in Turbulence 2 (Eds: H.-H. Fernholz, H. E. Fiedler), Springer, Berlin 1989, 438442.
[130] H. Choi, Ph.D. Thesis, Stanford University, CA 1993.
[131] J.-J. Wang, S.-l. Lan, G. Chen, Fluid Dyn. Res. 2000, 27 (4),
217229. DOI: 10.1016/S0169-5983(00)00009-5
[132] H. Ma, Q. Tian, H. Wu, J. Therm. Sci. 2005, 14 (3), 193197.
DOI: 10.1007/s11630-005-0001-7
[133] A. V. Boiko, K. H. Jung, H. H. Chun, I. Lee, J. Mech. Sci.
Technol. 2007, 21 (1), 196206. DOI: 10.1007/bf03161725
[134] B. Bhushan, Philos. Trans.: Math., Phys. Eng. Sci. 2009,
367 (1893), 14451486. DOI: 10.1098/rsta.2009.0011
[135] B. Dean, B. Bhushan, Philos. Trans.: Math., Phys. Eng. Sci.
2010, 368 (1929), 47754806. DOI: 10.1098/rsta.2010.0201
[136] G. D. Bixler, B. Bhushan, Adv. Funct. Mater. 2013, 23 (36),
45074528. DOI: 10.1002/adfm.201203683
[137] P. Ball, Nature 1999, 400 (6744), 507509. DOI: 10.1038/
22883
[138] W. E. Reif, Neues Jahrb. Geol. Palaeontol. 1982, 164,
172183.
[139] W.-E. Reif, A. Dinkelacker, Neues Jahrb. Geol. Palaeontol.
1982, 16, 184187.
[140] J. J. Videler, Symp. Soc. Exp. Biol. 1995, 49, 120.
[141] F. E. Fish, Bioinspiration Biomimetics 2006, 1 (2), R1725.
DOI: 10.1088/1748-3182/1/2/R01
[142] M. J. Walsh, NASA Langley Symposium on Aerodynamics,
Hampton, VA, April 1985.
[143] X. Xu, V. P. Carey, J. Thermophys. Heat Transfer 1990, 4 (4),
512520. DOI: 10.2514/3.215
[144] D. Moalem-Maron, R. Semiat, S. Sideman, Desalination 1980,
34 (3), 289309. DOI: 10.1016/S0011-9164(00)88595-1
[145] S. Nishimoto, B. Bhushan, RSC Adv. 2013, 3 (3), 671. DOI:
10.1039/c2ra21260a
[146] K. Liu, L. Jiang, Ann. Rev. Mater. Res. 2012, 42 (1), 231263.
DOI: 10.1146/annurev-matsci-070511-155046
[147] G. F. Naterer, P. S. Glockner, D. Thiele, S. Chomokovski,
G. Venn, G. Richardson, J. Micromech. Microeng. 2005,
15 (3), 501513. DOI: 10.1088/0960-1317/15/3/010
[148] L. James, K. Paul, R. Richard, in 38th Aerospace Sciences
Meeting and Exhibit, American Institute of Aeronautics and
Astronautics, Washington, DC 2000.
[149] F. Klocke, B. Feldhaus, S. Mader, Prod. Eng. 2007, 1 (3), 233
237. DOI: 10.1007/s11740-007-0031-y
[150] K. Wang, B. Song, G. Pan, Mech. Eng. 2005, 27 (2), 1820.
DOI: 10.6052/1000-0992-2003-221
[151] C. Lietmeyer, B. Denkena, T. Krawczyk, R. Kling, L. Overmeyer, B. Wojakowski, E. Reithmeier, R. Scheuer, T. Vynnyk,
J. R. Seume, J. Turbomach. 2013, 135 (4), 041008041008.
DOI: 10.1115/1.4007590
[152] T. Ninnemann, W. F. Ng, J. Fluids Eng. 2004, 126 (4),
642649. DOI: 10.1115/1.1667883
[153] B. Denkena, J. Kohler, T. Krawczyk, in New Production
Technologies in Aerospace Industry (Ed: B. Denkena), Springer, Berlin 2014, 6974.

www.ChemBioEngRev.de

[154] M. Worgull, in Hot Embossing 1 (Ed: J. J. Ramsden), William


Andrew Publishing, Boston 2009, 283306, Ch. 9.
[155] S.-W. Chen, J.-C. Hsieh, C.-T. Chou, H.-H. Lin, S.-C. Shen,
M.-J. Tsai, Sens. Actuators, A 2007, 139 (12), 7887. DOI:
10.1016/j.sna.2007.03.009
[156] V. D. Nguyen, J. Dickinson, Y. Jean, Y. Chalifour, A. Smaili,
A. Page, F. Paquet, in Turbulence Control by Passive Means,
Springer, Berlin 1990, 159172.
[157] B. Bhushan, in Biomimetics (Ed: B. Bhushan), Springer,
Berlin 2012, 227265, Ch. 10.
[158] L. P. Chamorro, R. E. A. Arndt, F. Sotiropoulos, Renewable
Energy 2013, 50 (0), 10951105. DOI: 10.1016/j.renene.
2012.09.001
[159] L. Reidy, G. Anderson, in 26th Aerospace Sciences Meeting,
American Institute of Aeronautics and Astronautics,
Washington, DC 1988.
[160] L. W. Reidy, Flat Plate Reduction in a Water Tunnel Using
Riblets, Naval Ocean Systems Center (NOSC), San Diego,
CA 1987.
[161] A. Gyr, E. Coustols, J. Cousteix, in Structure of Turbulence
and Drag Reduction, Springer, Berlin 1990, 577584.
[162] M. C. Gillcrist, L. W. Reidy, Drag and Noise Measurements
on an Underwater Vehicle with a Riblet Surface Coating,
Naval Ocean Systems Center (NOSC), San Diego, CA 1989.
[163] E. Coustols, in 2nd Shear Flow Conf., American Institute of
Aeronautics and Astronautics, Washington, DC 1989.
[164] J. J. Rohr, G. W. Andersen, L. W. Reidy, E. W. Hendricks,
Exp. Fluids 1992, 13 (6), 361368. DOI: 10.1007/
BF00223243
[165] G. W. Anderson, J. J. Rohr, S. D. Stanley, J. Fluids Eng. 1993,
115 (2), 213221. DOI: 10.1115/1.2910126
[166] W. Kyle, F. Saeed, in 14th Applied Aerodynamics Conf.,
American Institute of Aeronautics and Astronautics, Washington, DC 1996.
[167] S. Sundaram, P. Viswanath, N. Subaschandar, AIAA J. 1999,
37 (7), 851856. DOI: 10.2514/2.7533
[168] M. J. Walsh, I. W. L. Sellers, C. B. McGinley, J. Aircr. 1989,
26 (6), 570575. DOI: 10.2514/3.45804
[169] J. D. McLean, D. N. George-Falvey, P. P. Sullivan, Int. Conf.
on Turbulent Drag Reduction by Passive Means, London,
September 1987.
[170] Y. Furuya, I. Nakamura, M. Miyata, Y. Yama, Bull. JSME
1977, 20 (141), 315322. DOI: 10.1299/jsme1958.20.315
[171] M. J. Walsh, L. M. Weinstein, AIAA J. 1979, 17 (7), 770771.
DOI: 10.2514/3.61216
[172] M. J. Walsh, AIAA J. 1983, 21 (4), 485486. DOI: 10.2514/
3.60126
[173] G. V. Enyutin, Y. A. Lashkov, N. V. Samoilova, I. V. Fadeev,
E. A. Shumilkina, Fluid Dyn. 1987, 22 (2), 284289. DOI:
10.1007/bf01052264
[174] G. V. Enyutin, Y. A. Lashkov, N. V. Samoilova, I. V. Fadeev,
E. A. Shumilkina, Fluid Dyn. 1991, 26 (1), 3135. DOI:
10.1007/bf01050109
[175] S. P. Wilkinson, B. S. Lazos, in Turbulence Management and
Relaminarisation (Eds: H. W. Liepmann, R. Narasimha),
Springer, Berlin 1988, 121131.
[176] K.-S. Choi, in Turbulence Management and Relaminarisation (Eds: H. W. Liepmann, R. Narasimha), Springer, Berlin
1988, 149160.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

18

These are not the final page numbers! &&

[177] K. S. Choi, D. M. Orchard, Exp. Therm. Fluid Sci. 1997,


15 (2), 109124. DOI: 10.1016/s0894-1777(97)00047-2
[178] B. Frohnapfel, J. Jovanovic, A. Delgado, J. Fluid Mech. 2007,
590, 107116. DOI: 10.1017/s0022112007008221
[179] J. Jovanovic, B. Frohnapfel, A. Delgado, in Turbulence and
Interactions, Springer, Berlin 2010, 191197.
[180] S.-J. Lee, Y.-S. Choi, J. Turbul. 2008, 9 (23), 115. DOI:
10.1080/14685240802251517
[181] E. V. Bacher, C. Smith, in Shear Flow Control Conf., American Institute of Aeronautics and Astronautics, Washington,
DC 1985.
[182] C. J. A. Pulles, K. Krishna Prasad, F. T. M. Nieuwstadt, Appl.
Sci. Res. 1989, 46 (3), 197208. DOI: 10.1007/BF00404817
[183] M. J. Walsh, J. Aircr. 1990, 27 (6), 572573. DOI: 10.2514/
3.25323
[184] D. Neumann, A. Dinkelacker, Appl. Sci. Res. 1991, 48 (1),
105114. DOI: 10.1007/bf01998668
[185] K. Parker, A. T. Sayers, Proc. Inst. Mech. Eng., Part C 1999,
213 (8), 775785. DOI: 10.1243/0954406991522392
[186] R. Laurel, A. Greg, in 26th Aerospace Sciences Meeting,
American Institute of Aeronautics and Astronautics, Washington, DC 1988.
[187] K. N. Liu, C. Christodoulou, O. Riccius, D. D. Joseph, in
Structure of Turbulence and Drag Reduction, Springer, Berlin
1990, 545551.
[188] B. Dean, B. Bhushan, Appl. Surf. Sci. 2012, 258 (8),
39363947. DOI: 10.1016/j.apsusc.2011.12.067

www.ChemBioEngRev.de

[189] D. W. Bechert, G. Hoppe, J. G. T. van der Hoeven, R. Makris,


Exp. Fluids 1992, 12 (45), 251260. DOI: 10.1007/
BF00187303
[190] R. Gruneberger, W. Hage, Exp. Fluids 2011, 50 (2), 363373.
DOI: 10.1007/s00348-010-0936-7
[191] C. C. Buttner, U. Schulz, Smart Mater. Struct. 2011, 20 (9),
094016. DOI: 10.1088/0964-1726/20/9/094016
[192] J. Rohr, G. W. Anderson, L. W. Reidy, Drag Reduct. Fluid
Flows 1989, 4352.
[193] C. Christodoulou, K. N. Liu, D. D. Joseph, Phys. Fluids A
1991, 3 (5), 995. DOI: 10.1063/1.857980.
[194] H.-W. Bewersdorff, H. Thiel, Appl. Sci. Res. 1993, 19 (34),
347368. DOI: 10.1007/978-94-011-1701-2_10
[195] E. Koury, P. S. Virk, Appl. Sci. Res. 1995, 54 (4), 323347.
DOI: 10.1007/BF00863517
[196] H. Chen, X. Zhang, D. Che, D. Zhang, X. Li, Y. Li, Adv.
Mech. Eng. 2014, 2014, 18. DOI: 10.1155/2014/425701
[197] D.-Y. Zhao, Z.-P. Huang, M.-J. Wang, T. Wang, Y. Jin,
J. Mater. Process. Technol. 2012, 212 (1), 198202. DOI:
10.1016/j.jmatprotec.2011.09.002
[198] D. Zhang, Y. Li, X. Han, X. Li, H. Chen, Chin. Sci. Bull. 2011,
56 (9), 938944. DOI: 10.1007/s11434-010-4163-7
[199] D.-Y. Zhang, Y.-H. Luo, X. Li, H.-W. Chen, J. Hydrodyn., Ser.
B 2011, 23 (2), 204211. DOI: 10.1016/s1001-6058(10)
60105-9

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

19

These are not the final page numbers! &&

In fluid dynamics applications, a


substantial amount of energy is applied
to overcome drag forces. This paper
reviews passive drag reduction
techniques through bio-inspired riblets
that have shown advantages in
manipulating the turbulence boundary
layer. Different geometries and types
are discussed as well as the latest
developments in the technology presented.

www.ChemBioEngRev.de

Bio-Inspired Passive Drag Reduction


Techniques: A Review
H. A. Abdulabari*, H. D. Mahammad,
Z. B. Y. Hassan
ChemBioEng Rev. 2015, 2 (3),
XXX L XXX
DOI: 10.1002/cben.201400033

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemBioEng Rev 2015, 2, No. 3, 120

20

These are not the final page numbers! &&

You might also like