Anda di halaman 1dari 50

History of Aerodynamic Car Design

While scientists have more or less been aware of what it takes to create aerodynamic shapes for a
long time, it took a while for those principles to be applied to automobile design.

There was nothing aerodynamic about the earliest cars. Take a look at Ford's seminal Model T --
it looks more like a horse carriage minus the horses -- a very boxy design, indeed. Many of these
early cars didn't need to worry about aerodynamics because they were relatively slow. However,
some racing cars of the early 1900s incorporated tapering and aerodynamic features to one
degree or another.

In 1921, German inventor Edmund Rumpler created the Rumpler-Tropfenauto, which translates
into "tear-drop car." Based on the most aerodynamic shape in nature, the teardrop, it had a Cd of
just .27, but its unique looks never caught on with the public. Only about 100 were made [source:
Price].

On the American side, one of the biggest leaps ahead in aerodynamic design came in the 1930s
with the Chrysler Airflow. Inspired by birds in flight, the Airflow was one of the first cars
designed with aerodynamics in mind. Though it used some unique construction techniques and
had a nearly 50-50-weight distribution (equal weight distribution between the front and rear
axles for improved handling), a Great Depression-weary public never fell in love with its
unconventional looks, and the car was considered a flop. Still, its streamlined design was far
ahead of its time.

As the 1950s and '60s came about, some of the biggest advancements in automobile
aerodynamics came from racing. Originally, engineers experimented with different designs,
knowing that streamlined shapes could help their cars go faster and handle better at high speeds.
That eventually evolved into a very precise science of crafting the most aerodynamic race car
possible. Front and rear spoilers, shovel-shaped noses, and aero kits became more and more
common to keep air flowing over the top of the car and to create necessary downforce on the
front and rear wheels [source: Formula 1 Network].

On the consumer side, companies like Lotus, Citroën and Porsche developed some very
streamlined designs, but these were mostly applied to high-performance sports cars and not
everyday vehicles for the common driver. That began to change in the 1980s with the Audi 100,
a passenger sedan with a then-unheard-of Cd of .30. Today, nearly all cars are designed with
aerodynamics in mind in some way [source: Edgar].

What helped that change to occur? The answer: The wind tunnel. On the next page we'll explore
how the wind tunnel has become vital to automotive design.

Measuring Drag Using Wind Tunnels


To measure the aerodynamic effectiveness of a car in real time, engineers have borrowed a tool
from the aircraft industry -- the wind tunnel.

In essence, a wind tunnel is a massive tube with fans that produce airflow over an object inside.
This can be a car, an airplane, or anything else that engineers need to measure for air resistance.
From a room behind the tunnel, engineers study the way the air interacts with the object, the way
the air currents flow over the various surfaces.

The car or plane inside never moves, but the fans create wind at different speeds to simulate real-
world conditions. Sometimes a real car won't even be used -- designers often rely on exact scale
models of their vehicles to measure wind resistance. As wind moves over the car in the tunnel,
computers are used to calculate the drag coefficient (Cd).

Wind tunnels are really nothing new. They've been around since the late 1800s to measure
airflow over many early aircraft attempts. Even the Wright Brothers had one. After World War
II, racecar engineers seeking an edge over the competition began to use them to gauge the
effectiveness of their cars' aerodynamic equipment. That technology later made its way to
passenger cars and trucks.

However, in recent years, the big, multi-million-dollar wind tunnels are being used less and less.
Computer simulations are starting to replace wind tunnels as the best way to measure the
aerodynamics of a car or aircraft. In many cases, wind tunnels are mostly just called upon to
make sure the computer simulations are accurate [source: Day].

Many think that adding a spoiler on the back of a car is a great way to make it more
aerodynamic. In the next section, we'll examine different types of aerodynamic add-ons to
vehicles, and examine their roles in performance and providing better fuel mileage. The
Coefficient of Drag
We've just learned that the coefficient of drag (Cd) is a figure that measures the force of air
resistance on an object, such as a car. Now, imagine the force of air pushing against the car as it
moves down the road. At 70 miles per hour (112.7 kilometers per hour), there's four times more
force working against the car than at 35 miles per hour (56.3 kilometers per hour) [source:
Elliott-Sink].

The aerodynamic abilities of a car are measured using the vehicle's coefficient of drag.
Essentially, the lower the Cd, the more aerodynamic a car is, and the easier it can move through
the wall of air pushing against it.

Let's look at a few Cd numbers. Remember the boxy old Volvo cars of the 1970s and '80s? An
old Volvo 960 sedan achieves a Cd of .36. The newer Volvos are much more sleek and curvy,
and an S80 sedan achieves a Cd of .28 [source: Elliott-Sink]. This proves something that you
may have been able to guess already -- smoother, more streamlined shapes are more
aerodynamic than boxy ones. Why is that exactly?
Let's look at the most aerodynamic thing in nature -- a teardrop. The teardrop is smooth and
round on all sides and tapers off at the top. Air flows around it smoothly as it falls to the ground.
It's the same with cars -- smooth, rounded surfaces allow the air to flow in a stream over the
vehicle, reducing the "push" of air against the body.

Today, most cars achieve a Cd of about .30. SUVs, which tend to be more boxy than cars
because they're larger, accommodate more people, and often need bigger grilles to help cool the
engine down, have a Cd of anywhere from .30 to .40 or more. Pickup trucks -- a purposefully
boxy design -- typically get around .40 [source: Siuru].

Many have questioned the "unique" looks of the Toyota Prius hybrid, but it has an extremely
aerodynamic shape for a good reason. Among other efficient characteristics, its Cd of .26 helps it
achieve very high mileage. In fact, reducing the Cd of a car by just 0.01 can result in a 0.2 miles
per gallon (.09 kilometers per liter) increase in fuel economy [source: Siuru].

On the next page, we'll examine the history of aerodynamic design.

Aerodynamic Add-ons

There's more to aerodynamics than just drag -- there are other factors called lift and downforce,
too. Lift is the force that opposes the weight of an object and raises it into the air and keeps it
there. Downforce is the opposite of lift -- the force that presses an object in the direction of the
ground [source: NASA].

You may think that the drag coefficient on a Formula One racecar would be very low -- a super-
aerodynamic car is faster, right? Not in this case. A typical F1 car has a Cd of about .70.

Why is this type of racecar able to drive at speeds of more than 200 miles an hour (321.9
kilometers per hour), yet not as aerodynamic as you might have guessed? That's because
Formula One cars are built to generate as much downforce as possible. At the speeds they're
traveling, and with their extremely light weight, these cars actually begin to experience lift at
some speeds -- physics forces them to take off like an airplane. Obviously, cars aren't intended to
fly through the air, and if a car goes airborne it could mean a devastating crash. For this reason,
downforce must be maximized to keep the car on the ground at high speeds, and this means a
high Cd is required.

Formula One cars achieve this by using wings or spoilers mounted onto the front and rear of the
vehicle. These wings channel the flow into currents of air that press the car to the ground --
better known as downforce. This maximizes cornering speed, but it has to be carefully balanced
with lift to also allow the car the appropriate amount of straight-line speed [source: Smith].

Lots of production cars include aerodynamic add-ons to generate downforce. While the Nissan
GT-R supercar has been somewhat criticized in the automotive press for its looks, the entire
body is designed to channel air over the car and back through the oval-shaped rear spoiler,
generating plenty of downforce. Ferrari's 599 GTB Fiorano has flying buttress B-pillars designed
to channel air to the rear as well -- these help to reduce drag [source: Classic Driver].

But you see plenty of spoilers and wings on everyday cars, like Honda and Toyota sedans. Do
those really add an aerodynamic benefit to a car? In some cases, it can add a little high-speed
stability. For example, the original Audi TT didn't have a spoiler on its rear decklid, but Audi
added one after its rounded body was found to create too much lift and may have been a factor in
a few wrecks [source: Edgar].

In most cases, however, bolting a big spoiler on the back of an ordinary car isn't going to aid in
performance, speed, or handling a whole lot -- if at all. In some cases, it could even create more
understeer, or reluctance to corner. However, if you think that giant spoiler looks great on the
trunk of your Honda Civic, don't let anyone tell you otherwise.

For more information about automotive aerodynamics and other related topics, breeze on over to
the next page and follow the links.

Sources

 Classic Driver. "The Ferrari 599 GTB Fiorano." (March 9, 2009)


http://www.classicdriver.com/uk/magazine/3300.asp?id=12863
 Day, Dwayne A. "Advanced Wind Tunnels." U.S. Centennial of Flight Commission.
(March 9, 2009)
http://www.centennialofflight.gov/essay/Evolution_of_Technology/advanced_wind_tunn
els/Tech36.htm
 Edgar, Julian. "Car Aerodynamics Have Stalled." Auto Speed. (March 9, 2009)
http://autospeed.com/cms/A_2978/article.html
 Elliott-Sink, Sue. "Improving Aerodynamics to Boost Fuel Economy." Edmunds.com.
May 2, 2006. (March 9, 2009)
http://www.edmunds.com/advice/fueleconomy/articles/106954/article.html
 Formula 1 Network. "Williams F1 - History of Aerodynamics: Evolution of
aerodynamics." (March 9, 2009) http://www.f1network.net/main/s107/st22394.htm
 NASA. "Beginner's Guide to Aerodynamics." July 11, 2008. (March 9, 2009)
http://www.grc.nasa.gov/WWW/K-12/airplane/bga.html
 NASA. "The Drag Coefficient." July 11, 2008. (March 9, 2009)
 http://www.grc.nasa.gov/WWW/K-12/airplane/dragco.html
 Price, Ryan Lee. "Cheating Wind - Aerodynamic Tech and Buyers Guide: The Art Of
Aerodynamics And The Automobile." European Car Magazine. (March 9, 2009)
http://www.europeancarweb.com/tech/0610_ec_aerodynamics_tech_buyers_guide/index.
html
 Siuru, Bill. "5 Facts: Vehicle Aerodynamics." GreenCar.com. Oct. 13, 2008. (March 9,
2009) http://www.greencar.com/articles/5-facts-vehicle-aerodynamics.php
 Smith, Rich. "Formula 1 Aerodynamics." Symscape. May 21, 2007. (March 9, 2009)
http://www.symscape.com/blog/f1_aero
Automobile drag coefficient
From Wikipedia, the free encyclopedia

Tatra 77 maquette by Paul Jaray, 1933. This vehicle was the first serial-produced truly aerodynamically
designed automobile.

The drag coefficient is a common measure in automotive design as it pertains to aerodynamics.


Drag is a force that acts parallel and in the same direction as the airflow. The drag coefficient of
an automobile impacts the way the automobile passes through the surrounding air. When
automobile companies design a new vehicle they take into consideration the automobile drag
coefficient in addition to the other performance characteristics. Aerodynamic drag increases with
the square of speed; therefore it becomes critically important at higher speeds. Reducing the drag
coefficient in an automobile improves the performance of the vehicle as it pertains to speed and
fuel efficiency.[1] There are many different ways to reduce the drag of a vehicle. A common way
to measure the drag of the vehicle is through the drag area.

Contents
 1 Reducing drag
 2 Deletion
o 2.1 Roof rack
o 2.2 Mud flaps
o 2.3 Rear spoiler
o 2.4 Side mirrors
o 2.5 Radio antenna
o 2.6 Windshield wipers
 3 Fabrication
o 3.1 Wheel covers
o 3.2 Partial grille block
o 3.3 Under tray
o 3.4 Fender skirts
o 3.5 Modified front bumper
o 3.6 Boattails and Kammbacks
 4 Typical drag coefficients
 5 Drag area
 6 See also
 7 References
 8 External links
Reducing drag
The reduction of drag in road vehicles has led to increases in the top speed of the vehicle and the
vehicle's fuel efficiency, as well as many other performance characteristics, such as handling and
acceleration.[2] The two main factors that impact drag are the frontal area of the vehicle and the
drag coefficient. The drag coefficient is a unit-less value that denotes how much an object resists
movement through a fluid such as water or air. A potential complication of altering a vehicle's
aerodynamics is that it may cause the vehicle to get too much lift. Lift is an aerodynamic force
that flows perpendicular to the airflow around the body of the vehicle. Too much lift can cause
the vehicle to lose road traction which can be very unsafe.[3] Lowering the drag coefficient comes
from streamlining the exterior body of the vehicle. Streamlining the body requires assumptions
about the surrounding airspeed and characteristic use of the vehicle.

For high speed applications near or above the speed of sound, a Sears-Haack body,[4][5] is an
idealized shape that minimizes wave drag, which is the drag associated with supersonic shock
waves. This shape essentially consists of an elongated tube with pointed ends. This shape is seen
commonly in vehicles attempting to break speed and efficiency records, such as the North
American Eagle.[6]

Deletion
The deletion of parts on a vehicle is an easy way for designers and vehicle owners to reduce
parasitic and frontal drag of the vehicle with little cost and effort. Deletion can be as simple as
removing an aftermarket part, or part that has been installed on the vehicle after production, or
having to modify and remove an OEM part, meaning any part of the vehicle that was originally
manufactured on the vehicle. Most production sports cars and high efficiency vehicles come
standard with many of these deletions in order to be competitive in the automotive and race
market, while others choose to keep these drag-increasing aspects of the vehicle for their visual
aspects, or to fit the typical uses of their customer base.[7]

Roof rack

A roof rack is a common trait on many SUV and station wagon vehicles. While roof racks are
very useful in carrying extra storage on a vehicle, they also increase the frontal area of the
vehicle and increase the drag coefficient. This is because as the air flows over the top of the
vehicle, following the smooth lines of the hood and windshield, then collides with the roof rack
and causes turbulence. The removal of this part has led to increases in fuel efficiency in several
studies.[8]

Mud flaps

Mudflaps are now rarely specified as standard on production cars as they interfere with the clean
airflow around the vehicle. For larger vehicles such as trucks, mud flaps are still important for
their control of spray, and in 2010 a new version of the mud flap was introduced that has been
shown to create significantly less aerodynamic drag than standard mud flaps.[9]
Rear spoiler

A rear spoiler usually comes standard in most sports vehicles and resembles the shape of a raised
wing in the rear of the vehicle. The main purpose of a rear spoiler in a vehicle's design is to
reduce lift, thereby increasing stability at higher speeds. In order to achieve the lowest possible
drag, air must flow around the streamlined body of the vehicle without coming into contact with
any areas of possible turbulence. A rear spoiler design that stands off the rear deck lid will
increase downforce, reducing lift at high speeds while incurring a drag penalty. Flat spoilers,
possibly angled slightly downward may reduce turbulence and thereby reduce the coefficient of
drag.[10] Some cars now feature automatically adjustable rear spoilers, so at lower speed the
effect on drag is reduced when the benefits of reduced lift are not required.

Side mirrors

Side mirrors both increase the frontal area of the vehicle and increase the coefficient of drag
since they protrude from the side of the vehicle.[11][12] In order to decrease the impact that side
mirrors have on the drag of the vehicle the side mirrors can be replaced with smaller mirrors or
mirrors with a different shape. Several concept cars of the 2010s are replacing mirrors with tiny
cameras[13] but this option is not common for production cars because most countries require side
mirrors.

Radio antenna

While they do not have the biggest impact on the drag coefficient due to their small size, radio
antennas commonly found protruding from the front of the vehicle can be relocated and changed
in design to rid the car of this added drag. The most common replacement for the standard car
antenna is the shark fin antenna found in most high efficiency vehicles.[14]

Windshield wipers

The effect that windshield wipers have on a vehicles airflow varies between vehicles; however,
they are often omitted from race vehicles and high efficiency concepts in order to maintain the
smallest possible coefficient of drag. A much more common option is to replace the windshield
wipers with lower profile wipers, or to only remove the windshield wiper on the passenger side
of the vehicle, and even to fabricate a deflector to deflect the air up and over the wipers.[15]

Another alternative is to equip the vehicle with a single wiper placed in the centre of the
windshield, allowing it to cover both sides of the windshield. This mitigates the amount of drag
by decreasing the surface area of the blade. While the application of a single wiper blade is
useful for performance level vehicles, a common street car would see only marginal
improvements in both fuel efficiency and acceleration/speed.

Fabrication
The application of new parts and concepts onto the vehicle design are easier to include when in
the design stage of a vehicle, rather than in aftermarket (automotive) parts, however, the
fabrication of these parts assists in the streamlining of the vehicle and can help greatly reduce the
drag of the vehicle. Most vehicles with very low drag coefficients, such as race cars and high
efficiency concept cars, apply these ideas to their design.[16]

Wheel covers

When air flows around the wheel wells it gets disturbed by the rims of the vehicles, and forms an
area of turbulence around the wheel. In order for the air to flow smoother around the wheel well
smooth wheel covers are often applied. Smooth wheel covers are hub caps with no holes in them
for air to pass through. This design reduces drag, however, it may cause the brakes to heat up
quicker because the covers prevent airflow around the brake system. This is why this
modification is more commonly seen with high efficiency vehicles, rather than sports cars or
racing vehicles.[17]

Partial grille block

The front grille of a vehicle is used to direct air directly into the engine compartment. In a
streamlined design the air flows around the vehicle rather than through; however, the grille of a
vehicle redirects airflow from around the vehicle to through the vehicle, which then increases the
drag. In order to reduce this impact a grille block is often used. A grille block covers up a
portion, or the entirety of the front grille of a vehicle. In most high efficiency models or vehicles
with low drag coefficients there will be a very small grille already built into the design, therefore
a grille block is unneeded. The grille in most production vehicles is built generally to have as
much air flowing into the engine in order to keep it from overheating. But most commonly there
is too much airflow into the engine, preventing it from warming up in a timely manner, so a
grille block is used to increase engine performance and reduce the vehicle's drag.[18]

Under tray

The underside of a vehicle often traps air in various places and adds turbulence around the
vehicle. In most racing vehicles this is eliminated by covering the entire underside of the vehicle
in what is called an under tray. This tray prevents any air from becoming trapped under the
vehicle and reduces drag.[17]

Fender skirts

Fender skirts are often made as extensions of the body panels of the vehicles and cover the entire
wheel wells. Much like smooth wheel covers this modification reduces the drag of the vehicle by
preventing any air from becoming trapped in the wheel well and assists in streamlining the body
of the vehicle. Fender skirts are more commonly found on the rear wheel wells of a vehicle
because the tires do not turn and the design is much simpler. This is commonly seen in vehicles
such as the first generation Honda Insight. Front fender skirts have the same effect on reducing
drag as the rear wheel skirts, but must be further offset from the body in order to compensate for
the tire sticking out from the body of the vehicle as turns are made.[17]
Modified front bumper

The front bumper is the first part of the vehicle that the air must flow around. Therefore, it plays
a crucial role in reducing drag. In order to preserve the teardrop shape of the vehicle a front air
dam is often used. A front air dam extends from the very front of the vehicle down to the lowest
part of the vehicle. It does this to direct airflow around the vehicle rather than through it.
Contoured deflectors, or tire spats, are often made as part of the front bumper in order to direct
airflow around the tire without having any increase to the outward flow.

Boattails and Kammbacks

A boattail can greatly reduce a vehicle's total drag. Boattails create a teardrop shape that will
give the vehicle a more streamlined profile, reducing the occurrence of drag inducing flow
separation.[19] A kammback is a truncated boattail. It is created as an extension of the rear of the
vehicle, moving the rear backward at a slight angle toward the bumper of the car. This can
reduce drag as well but a boattail would reduce the vehicles drag more. Nonetheless, for practical
and style reasons, a kammback is more commonly seen in racing, high efficiency vehicles, and
trucking.[20]

Typical drag coefficients


This section needs additional citations for verification. Please help improve this article by
adding citations to reliable sources. Unsourced material may be challenged and removed.
(November 2013)

The average modern automobile achieves a drag coefficient of between 0.30 and 0.35. SUVs,
with their typically boxy shapes, typically achieve a Cd=0.35–0.45. The drag coefficient of a
vehicle is affected by the shape of body of the vehicle. Various other characteristics affect the
coefficient of drag as well, and are taken into account in these examples. Some sports cars have a
surprisingly high drag coefficient, but this is to compensate for the amount of lift the vehicle
generates, while others use aerodynamics to their advantage to gain speed and have much lower
coefficients of drag.

Some examples of Cd follow. Figures given are generally for the basic model. Some "high
performance" models may actually have higher drag, due to wider tires, extra spoilers and larger
cooling systems as many basic / low power models have half size radiators with the remaining
area blanked off to reduce cooling and engine bay drag. The Opel/Vauxhall Calibra 8V being a
notable example.

Production cars Production cars (continued) Concept/experimental cars

Calendar Calendar Calenda


Cd Automobile Cd Automobile Cd Automobile
Year Year r Year
typical values 1975– 0.27 Avion[114] 1986
0.32 AMC Pacer
for a Formula 1980
0.7 One car Alfa Romeo
0.26 1952
to (downforce 0.32 Buick Riviera 1995 Disco Volante
1.1 settings
BMW M3 0.25 Dymaxion Car 1933
change for 0.32 2005
Coupe
each circuit)
SmILE (an
Dodge 0.25 experimental 1996
0.74 Legends car 0.32 1995
Avenger car)
Caterham
0.7 Ferrari BMW Vision
Seven 0.32 2008
California 0.22 EfficientDynami 2009
0.65 cs Concept
1957– Chrysler 2011-
to Lotus Seven 0.32
1972 300C 2014 Citroën ECO
0.75 0.22 1981[115]
2000 Concept
1985-
a typical 0.32 Fiat Croma
0.6 + 1996 0.20 Loremo Concept 2006
truck
1992- Opel Eco
0.57 Hummer H2 2003 0.32 Ford Taurus
1995[62] 0.20 Speedster 2003
Mercedes Concept[116]
0.54 Geo Metro 1995-
Benz G-Class 0.32
(Sedan) 1997[34] Alfa Romeo
0.19 1954
Volkswagen B.A.T. 7 Concept
1980- 2005-
0.51 Westfalia 0.32 Seat Leon
1991 2011 Dodge Intrepid
Camper 0.19 1995
ESX Concept
Honda
0.51 Citroën 2CV 1948
0.32 Accord 2002 General Motors
0.19 1992
0.48 Rover Mini 1998 (Coupe) Ultralite

Volkswagen Honda Ascot Mercedes-Benz


1992-
Beetle 0.32 Innova Bionic
0.48 1938 1996 0.19 2005
(original (Sedan) Concept[117](base
design)[21][22] d on the boxfish)
Honda Civic 1992-
0.32
Volkswagen (Coupe) 1995[34] Chrysler Ghia
1979– 0.170 1955[118]
0.48 Cabriolet Dart
1993
0.32 Honda Civic
(Rabbit 1996-
(Hatchback
Convertible)[2 DX) 2000[63] Pininfarina Fiat
3]
0.17 124 concept 1978
Honda Civic 1996- (Morelli shape)
0.32
Lancia (Sedan EX) 2000[64]
0.47 1937
Aprilia[24] Daihatsu UFE-III
Honda 0.168 2005[119]
Concept
Ford 0.32 Insight 2009-
0.46 Mustang 1979 Hybrid [65] General Motors
(coupe) 0.16 Precept Concept 2000
0.32 Honda NSX 1990 (5 seats)[120]
Dodge Viper
0.45 1996 Hyundai
RT/10 0.32 2012 Edison2 Very
Veloster Light Car,
Range Rover 0.16 2010
0.45 1990 Automotive X
Classic Jaguar XJ
0.32 2006 Prize winner[121]
(X350)
Volkswagen 1980- Volkswagen 1-
0.44 Koenigsegg 0.159 2002
Vanagon 1991 0.32 2006 litre car Concept
CCX
Ford Li-ion Motors
0.44 Mustang 1979 Mazdaspeed
0.32 2007 Wave II,
(fastback) 3 0.157 2010
Automotive X
Prize winner[122]
0.44 Peugeot 305 1978 0.32 McLaren F1 1992
0.15 Schlörwagen[123] 1939
0.44 Peugeot 504 1968 Mercedes-
Benz 190E
0.32 Aptera 2 Series
0.44 Toyota Truck 1990 2.5-16/2.3- 0.15 2011
2e Prototype
16
0.43 TVR 3000S 1978-79
0.149
Nissan 240SX 1995-
0.32 - Urbee 2 [124] 2013
Duple 425 Coupe 1998[66]
0.150
0.42 coach
(named for its 1985 0.32 Nissan 300ZX 1989
5 JCB Dieselmax
low Cd by coach
standards) 0.147 land speed 2006
Nissan 1998-
0.32 record holder
Altima 2001[67]
Lamborghini
0.42 1974 Fiat Turbina
Countach Nissan 0.14 1954
0.32 1997 Concept
Maxima
0.42 Plymouth 1994
0.137 Ford Probe V 1985
Duster Oldsmobile 1995- Concept
0.32
Aurora 1999
Triumph Sunraycer, solar
0.42 1971 [68]
0.125 1987
Spitfire Mk IV 0.32 Opel Astra J 2009 race car

Smart Porsche 997 2008– Reflex 1000,


0.41 2003 0.32 0.12 1996[125]
Roadster GT2 2013 solar cycle

Volvo 740 0.32 Peugeot 406 1995 Summers


0.41 1982
(sedan) Brothers
0.32 Peugeot 806 1994 0.117 Goldenrod 1965
0.40 Subaru 1997- Bonneville race
5 Forester 2002[25] Saab Sonett 1966-
0.32 car
II 1969 [69]
Mercury 1983- Fortis Saxonia
0.40 0.32 Scion xB 2008
Cougar 1986 (Shell Eco-
0.08 2007
marathon)
Chevrolet 1995- 0.32 Suzuki Swift 1991
0.40 Concept
Astro 2005[26]
1948-
0.32 Tatra 600 PAC-Car II (Shell
0.40 Ariel Atom 2002 [27] 1952[70]
0.075 Eco-marathon) 2005
0.32 Toyota Celica 1994 Concept
Ford Escape
0.40 2005
Hybrid
2000- Alérion
0.32 Toyota Celica Supermileage
Nissan 2005[71]
0.072 (Shell Eco- 2013
0.40 Skyline GT-R 1989
Toyota Supra marathon)
R32
(N/A with Prototype
1976- 0.32 wing and 1993
0.40 Jaguar XJS turbo Nuna, World
1996 2001–
models) 0.07 Solar Challenge
2007
Chevrolet winner
0.39 2006 Toyota Supra
Tahoe 1987–
0.32 (with factory
1988
Dodge turbo wing)
0.39 2004
Durango
Toyota 1995-
[28]
0.32
0.39 Ford Aerostar 1995 Tercel Sedan 2000[72]

0.39 Ford Escort 5 1981- 0.32 Volkswagen 1991


Door 1984[29] Golf Mk3

Honda Volkswagen
0.39 1994-98 0.32 2006
Odyssey GTI Mk V

Triumph 0.32 Volvo V50 2004


0.39 1964
Spitfire
1996-
0.315 Saturn SL1
0.38 1999[73]
Nissan 280ZX 1978
5
0.31 Audi A4 B5 1995
0.38 Lexus GX 2003
2011–
0.31 Audi A5
0.38 Mazda Miata 1989 present

0.38 Fiat 500 1957 0.31 Audi A3 2014

Smart BMW 7-
0.31 2009
0.38 Roadster 2003 series
Coupé
Buick Park
0.31 1996
Subaru 2009- Avenue
0.38
Forester 2013[30]
0.31 Cadillac CTS 2004
VW
NewBeetle[31] Cadillac CTS-
0.31 2005
0.38 without wing 2003 V
or spoiler
0.31 Citroën AX 1986
0.39[32]
0.31 Citroën GS 1970
0.37 Ford Capri
1978
4 Mk III
0.31 Eagle Vision 1995
0.37
Ferrari F50 1996 0.31 Fiat Coupé 1995[74]
2
1988-
Ford Escort 0.31 Fiat Tipo
1995
0.37 Mk.III 1980
(Europe) 0.31 Ford Falcon 1995

0.37 BMW Z3 M 1999 0.31 Ford 1989-


coupe Thunderbird 1997[75]

Jaguar XJ Holden
0.37 0.31 1998
(X300/X308) Commodore

Renault Honda Civic 1992-


0.37 0.31
Twingo (Hatchback) 1995[34]

Volkswagen Honda Civic


0.37 2008 0.31 2006
Tiguan (Sedan)

Mercury 1987- 2008–


0.36 Infiniti G37
Cougar 1988 0.31 present[7
(Coupe) 6]

Alfa Romeo
0.36 1983[33]
33 Kia Rio
0.31 2001[77]
(Sedan)
Cadillac
0.36 Escalade 2008 Lamborghini
0.31 1990
hybrid Diablo

Cadillac Lexus LFA


0.36 1996
Fleetwood 0.31 (wing 2010
retracted)
Citroën CX
0.36 (named after 1974 1990–
0.31 Mazda MX-3
the term for Cd) 1996

0.36 Citroën DS 1955 1992–


0.31 Mazda MX-6
1997
Chrysler
0.36 1996
Sebring Mazda RX-7
0.31 1986
FC3S
Ferrari
0.36 1986
Testarossa Mazda RX-7
0.31 1993
FD R1(R2)
1997-
0.36 Ford Escort
2002[34] 0.31 Mazda RX-8 2004

Ford Mazda3
0.36 1999 0.31 2010
Mustang (Hatchback)
2001– Nissan Tiida /
0.36 Honda Civic 0.31 2004
2005 Versa

1949- 1994-
0.36 Nash Airflyte 0.31 Opel Tigra
1951 2000

0.36 Opel GT 1969 Pagani


0.31 2012[78]
Huayra
Subaru
0.36 Impreza 2010 0.31 Peugeot 307 2001
WRX[35]
0.31 Peugeot 405 1987
1996-
0.36 Saturn SW Porsche 997
2001[34] 0.31 2006
Turbo/GT3
0.36,
0.24 1936- 0.31 Renault 25 1984
Tatra 87
4 1950[36]
Saab Sonett 1970-
(1:5) 0.31
III 1974 [79]
Toyota Celica 1994-
0.36 0.31 Saturn SC2 2001
Convertible 1999[37]
0.31 Scion xA 2004
Volkswagen 1985-
0.36
Jetta 1992[38]
Toyota 1995–
0.31
Avalon 2000
0.35
NSU Ro 80 1967
5
Toyota 1998-
0.31
Corolla 2002[80]
Ford 1983-
0.35
Thunderbird 1988
1995-
0.31 Toyota Paseo
1999[81]
1986-
0.35 Aleko 2141
2002
0.31 Toyota RAV4 2006
Aston Martin
0.35 2004 Toyota
Vanquish
0.31 Camry 1992
(Sedan)
BMW M3
0.35 2005
Convertible
0.31 Toyota Supra 1993
(N/A;
BMW Z4 M without
0.35 2006
coupe factory wing)

DeltaWing[39] Volkswagen
0.31 1997
0.35 (endurance 2012 GTI Mk IV
racing car)
Volkswagen 2008-
0.31
Dodge Viper Golf Mk6 2012
0.35 1996
GTS
Volvo S40
Honda Del 1992– 0.31 2nd 2003
0.35
Sol 1997[34] generation

0.35 Jaguar XKR 2005 1988-


0.304 Ford Probe
1992[82]
0.35 Lexus GX 2010
Alfa Romeo
2003– 0.30 1988[83]
0.35 Lexus RX 164
2009
0.30 Audi 100 1983
0.35 MINI Cooper 2008
Fiat Uno 2nd 1989-
0.35 Nissan Cube 2009 0.30
gen. 2000

Renault Clio 1996-


0.35 2002[40]
(Mk 2) 0.30 Ford Taurus 1999[84][8
5]
SSC Ultimate 2007–
0.35
Aero present Ford Focus 2013–
0.30
ST present
Tesla
0.35 2008
Roadster[41] Honda 2003,
0.30 Accord 2005–
Mitsubishi i-
0.35 2011 Sedan 2007
MiEV
Honda CRX
Smart 0.30 1988
0.35 2008- DX/Si[86]
ForTwo[42]
0.30 Honda NSX 2002
0.35 Toyota MR-2 1998

0.30[citatio Honda 2005


n needed]
1991- Odyssey
0.35 Toyota Previa 1997[43][4
4] Hyundai
0.30 2006
Sonata
Toyota
0.35 2007 Koenigsegg
Sequoia 0.30 2006
CCX
Volvo 940
0.35 1990 Mitsubishi
(sedan) 0.30 2000
Eclipse
0.34 Toyota Celica
1982[45] 0.30 Nissan 180SX 1989
8 Supra (Mk 2)

Toyota Celica 0.30 Nissan 300ZX 1983


0.34
(Liftback 1982
2 Nissan 350Z
Model)
Coupe Base
2003–
Aston Martin 0.30 and
0.34 2004 2008
DB9 Enthusiast
models
Chevrolet
0.34 1994 Nissan 370Z
Caprice
Coupe
0.30 2009[87]
Chevrolet C6 2006– (0.29 with
0.34 sport package)
Corvette ZO6 2013

Chevrolet Renault 19
0.34 2008 0.30 1991
Tahoe hybrid 16V

Ferrari 360 0.30 Saab 92 1947


0.34 1999
Modena
0.30 Seat Leon 2012
0.34 Ferrari F40 1987
Toyota 2003-
0.30
Ferrari F430 Corolla 2008[88]
0.34 2004
F1
Toyota 2003–
0.30
0.34 Ford Puma 1997 Sienna 2009

0.34 Ford Sierra 1982 Volkswagen 1999–


0.30
Bora mk4 2005
Geo Metro 1995- Mercedes-
0.34
(Hatchback) 1997[34] Benz CLA 2013–
0.30
BlueEfficienc Present
Honda y Edition [89]
0.34 1988
Prelude
0.299 Cadillac ATS 2012 [90]
Mercedes-
0.34 Benz SL (Roof 2001 1990-
0.297 Fiat Tempra
Down) 1999

1993- 0.295 Ford Falcon 1998


0.34 Nissan Altima
1997[46]
Ford Focus
Nissan 0.295 Mk.III 2011
1999-
0.34 Skyline R34 hatchback [91]
2002[47]
GT-R
Toyota
0.291 2005
0.34 Peugeot 106 1991 Avalon

1991- Alfa Romeo


0.34 Saturn SL2 0.29 1992[92]
1995[48] 155

Subaru 2004-
0.29 Acura TL
0.34 Impreza WRX 2009[49] 2008
(4 Door)
0.29 Audi 80 1991
Subaru
1993- 2011–
0.34 Legacy 0.29 Audi A4
1999[50] present
Wagon

Toyota BMW 1-
1993- 0.29 Series (116i 2008
0.34 Corolla
1997[51] Sportshatch)
(Wagon)

Toyota Supra BMW 8- 1989-


0.29
(with factory 1989– Series 1999
0.34
3 piece turbo 1990
0.29 BMW i3 2013
wing)
Chevrolet
1984- 0.29 2005
0.34 Saab 9000 Corvette
1998 [52]
0.33 Chevrolet Chevrolet
1995
8 Camaro 0.29 Corvette C5 2002
ZO6
0.33 2005-
Seat Leon FR
4 2011 Daewoo
0.29 1990
Espero
1993-
0.33 Acura Integra
2001[53] Dodge
0.29 Charger 1969
2002- Daytona
0.33 Acura RSX
2006[54]
0.29 Eagle Talon 1990s
Alfa Romeo
0.33 Giulia 1962[55] 0.29 Ford Escape 2010[93]
(saloon)
Ford Focus C-
0.29 2003[94]
0.33 Audi A3 2006 Max

Dodge Honda 2003,


0.33 2006
Charger 0.29 Accord 2005–
Coupe 2007
Ford Crown
0.33 1992
Victoria Honda
2005,
0.29 Accord
Ford Escort 1998- 2007
0.33 Hybrid
ZX2 2003[56]
Honda CRX
0.33 Ford Fusion 2010[57] 0.29 1988
HF[86]

Honda Infiniti G35


0.33 2002 0.29 2008
Accord Sedan Sedan

Honda Civic 1988- 0.29 Lancia Dedra 1990


0.33
Hatchback 1991[34]
Lexus CT 2011–
0.33 0.29
Koenigsegg 200h present
to 2013
Agera (R)
0.37 0.29 Lexus LS 400 1990

Lamborghini 0.29 Lotus Elite 1958


0.33 2001
Murcielago
0.33 Lexus RX 2010 0.29 Lotus Europa 1966

Mazda RX-7 Mazda


0.33 1987 0.29 1995
FC3C Millenia

Nissan 200SX 1995- Mazda RX-7


0.33
Coupe 1998[58] 0.29 FC3S Aero 1986
Package
0.33 Peugeot 206 1998
Mazda RX-7
Dodge 0.29 1993
FD
Durango 2011–
0.33 (without roof- 0.29 Mazda3 2010
present
rack), ("HEAT"
model, 0.325) Mercedes-
2001–
0.29 Benz SL
0.33 Peugeot 309 1986 present
(Roof Up)
Renault
0.33 2004 Mercedes-
Modus
0.29 Benz C-Class 2001
Sportscoupe
Subaru
0.33 Impreza WRX 2004
Nissan 350Z
STi
Coupe Track 2007–
0.29
and Grand 2008
Subaru 2014-
0.33 Touring
Forester Current
2007–
0.33 Saturn SL2 1999[59] 0.29 Nissan Versa
2008
Toyota 1993-
0.33 Opel Calibra
Corolla 1997[34] 1989-
0.29 (16v / V6 /
1997
Turbo versions)
Toyota Supra
1989–
0.33 (without 0.29 Peugeot 208 2012
1990
wing)
2007–
0.32 Chevrolet 1989- 0.29 Peugeot 308
present
9 Corsica 2006[60]
2004-
2005 0.29 Peugeot 407
0.32 Cobalt SS 2011
4 Supercharged 2000-
0.29 Peugeot 607
2010
0.32 Toyota 2003-
1 Matrix 2008[61] Pontiac
Firebird
Trans Am
0.29 (with optional 1984
W62 Aero
Package and
N89 Turbo Cast
rims)

0.29 Porsche 918 2010

Porsche 2005–
0.29
Boxster present

Subaru SVX
(Without
0.29 1992
factory
spoiler)

Toyota 1996–
0.29
Camry 2001

2000-
0.29 Toyota Echo 2005[95][9
6]

0.29 Toyota Prius 2001

2006-
0.29 Toyota Yaris
2011

Volvo 850 T-
0.29 1995
5R sedan

0.29 Volvo C70 2000

Chrysler 1998–
0.288
Concorde 2001
Chevrolet 2005–
0.286
Corvette C6 present

Volkswagen 2008–
0.284
Passat CC present

Chevrolet
0.281 2010
Volt

Audi A2 1.4
0.28 2000
TDI

Mercedes- 1985-
0.28
Benz E-Class 1996

0.28 Citroën C4 2004

0.28 Citroën XM 1989

Fiat Croma 2005-


0.28
Nuova 2011

Honda Civic 2003-


0.28
Hybrid 2005[34]

Hyundai
0.28 2011
Elantra

Hyundai
Sonata 2011-
0.28
(0.25 for the 2013
Hybrid)

2006–
0.28 Lexus IS
present

0.28 Lexus LS400 1998

Luxgen5
0.28 2011
Sedan

0.28 Mazda3 2012


(Hatchback)

Mitsubishi
0.28 1995
Diamante

0.28 Nissan Leaf 2011

0.28 Opel Astra F 1991

Opel Omega
0.28 1986
A

Porsche 997
Carrera
(with optional
0.28 automatic 2004
spoiler, PDK
transmission
0.30)

Renault 25
0.28 1984
TS

Rumpler
0.28 Tropfenwage 1921
n

0.28 Saab 9-3 2003 [97]

Toyota
0.28 Camry / 2001
Lexus ES

Chevrolet
0.28 Cruze sedan 2015
[98]

Toyota Auris
0.277 2013
hatchback

0.275 Ford Fusion 2013[99]


Ford Focus
0.274 Mk.III sedan 2011
[91]

Audi A2 1.6
0.27 2003
FSI

Honda Civic
0.27 2006-
Hybrid

Hyundai
0.27 2009
Genesis

Infiniti G35
Coupe 2003–
0.27 (0.26 with
2007
"aero
package")

0.27 Lexus GS 2005

0.27 Mazda6 2009

Mazda6
(sedan and
0.27 2008
hatchback)[10
0]

Mercedes-
Benz W203
0.27 2001
C-Class
Sedan

Mercedes-
Benz S Class 2000–
0.27
(0.268 with 2005
Sport Package)

2007–
0.27 Nissan GT-R
2010
Opel Insignia
2008–
0.27 (0.26 in
present
EcoFlex trim)

0.27 Subaru BRZ 2013[101]

Toyota
0.27 Camry 2007
Hybrid

Tucker 48
0.27 1948
(Torpedo)

Volkswagen 2012–
0.27
Golf Mk7 present

Volkswagen
0.27 Passat B5 1997
(sedan)

BMW E90
0.26 2009[102]
(0.26-0.30)

0.26 BMW i8 2014

0.26 Jaguar XE[103] 2014

Lexus LS 430
2001–
0.26 (without air
2006
suspension)

0.26 Lexus LS 460 2006

Mazda3
0.26 2012[104]
(Sedan)

Mercedes-
Benz B-Class
0.26 2012-
+ eco-
package

0.26 Mercedes- 2002-


Benz E-Class 2009

Mercedes-
2006–
0.26 Benz W221
present
S-Class

2013–
0.26 Infiniti Q50
present

2011–
0.26 Nissan GT-R
present

Opel Calibra
0.26 (8 valve 1989
version)

2004–
0.26 Toyota Prius
2009

0.26 Citroen SM 1970

2011–
0.25 Peugeot 508
present

Lexus LS 430
2001–
0.25 (with air
2006
suspension)

Audi A2 1.2
0.25 2001
TDI

Honda 1999-
0.25
Insight 2006

Hyundai
0.25 Sonata 2013
Hybrid

0.25 Toyota Prius 2010

0.2455
Tatra 77 1933
(1:5)
Mercedes-
2014–
0.24 Benz S-
Present
Class[105]

Mercedes-
2014–
0.24 Benz C-
Present
Class[106]

Tesla Model
0.24 2012
S[107]

0.212
(according
1935[110]
to some
sources
Tatra T77A
[111] [112]
1:5)[108][10 [113]
9]

General
0.195 1996
Motors EV1

Volkswagen
0.189 2013
XL1

Drag area
While designers pay attention to the overall shape of the automobile, they also bear in mind that
reducing the frontal area of the shape helps reduce the drag. The combination of drag coefficient
and area - drag area - is represented as CdA (or CxA), a multiplication of the Cd value by the
area.

The term drag area derives from aerodynamics, where it is the product of some reference area
(such as cross-sectional area, total surface area, or similar) and the drag coefficient. In 2003, Car
and Driver magazine adopted this metric as a more intuitive way to compare the aerodynamic
efficiency of various automobiles.

The force required to overcome drag is:


As drag area CdA is the fundamental value that determines power required for a given cruise
speed it is a critical parameter for fuel consumption at a steady speed. This relation also allows
an estimation of the new top speed of a car with a tuned engine, estimated top speed = original
top speed x cube root (new power / original power). Or the power required for a target top speed,
power required = original power x (target speed / original speed)³.

Average full-size passenger cars have a drag area of roughly 8.50 sq ft (0.790 m2). Reported drag
areas range from the 1999 Honda Insight at 5.1 sq ft (0.47 m2) to the 2003 Hummer H2 at
26.5 sq ft (2.46 m2). The drag area of a bicycle is also in the range of 6.5–7.5 sq ft (0.60–
0.70 m2).[126]

Automobile examples of CdA[127]

CdA sqft CdA m2 Automobile model

2.50 sq ft 0.232 m2 1986 Twike[128]

2.69 sq ft 0.250 m2 2009 Loremo

3.00 sq ft 0.279 m2 2011 Volkswagen XL1

3.95 sq ft 0.367 m2 1996 GM EV1

5.00 sq ft 0.465 m2 2005 Mercedes-Benz Bionic[129]

5.10 sq ft 0.474 m2 1999 Honda Insight

5.40 sq ft 0.502 m2 1989 Opel Calibra

5.70 sq ft 0.530 m2 1985 Subaru Alcyone/XT/Vortex

5.71 sq ft 0.530 m2 1990 Honda CR-X Si

5.74 sq ft 0.533 m2 2002 Acura NSX

5.76 sq ft 0.535 m2 1968 Toyota 2000GT

5.80 sq ft 0.539 m2 1986 Toyota MR2

5.81 sq ft 0.540 m2 1989 Mitsubishi Eclipse GSX

5.86 sq ft 0.544 m2 2001 Audi A2 1.2 TDI 3L

5.88 sq ft 0.546 m2 1990 Nissan 240SX / 200SX / 180SX


Automobile examples of CdA[127]

CdA sqft CdA m2 Automobile model

5.92 sq ft 0.550 m2 1994 Porsche 911 Speedster

5.95 sq ft 0.553 m2 1990 Mazda RX7

6.00 sq ft 0.557 m2 1992 Subaru SVX

6.00 sq ft 0.557 m2 1970 Lamborghini Miura

6.08 sq ft 0.565 m2 2008 Nissan GTR

6.08 sq ft 0.565 m2 1989 Geo Metro [34]

6.13 sq ft 0.569 m2 1991 Acura NSX

6.17 sq ft 0.573 m2 1995 Lamborghini Diablo

6.19 sq ft 0.575 m2 1981 Citroën GSA X3 [130]

6.20 sq ft 0.576 m2 2012 Tesla Model S [131]

6.20 sq ft 0.576 m2 2014 Toyota Prius

6.24 sq ft 0.580 m2 2004 Toyota Prius

6.27 sq ft 0.583 m2 1986 Porsche 911 Carrera

6.27 sq ft 0.583 m2 1992 Chevrolet Corvette

6.35 sq ft 0.590 m2 1999 Lotus Elise

6.37 sq ft 0.592 m2 2000 Vauxhall VX220 N/A

6.40 sq ft 0.595 m2 1990 Lotus Esprit

6.41 sq ft 0.596 m2 2003 Smart Roadster Coupé

6.54 sq ft 0.608 m2 1991 Saturn Sports Coupe

6.57 sq ft 0.610 m2 1985 Chevrolet Corvette


Automobile examples of CdA[127]

CdA sqft CdA m2 Automobile model

6.63 sq ft 0.616 m2 2001 Audi A2

6.66 sq ft 0.619 m2 1996 Citroën Saxo

6.70 sq ft 0.622 m2 2014 Chevrolet Volt

6.77 sq ft 0.629 m2 1995 BMW M3

6.79 sq ft 0.631 m2 1993 Toyota Corolla DX

6.80 sq ft 0.632 m2 2007 BMW 335i Coupe

6.81 sq ft 0.633 m2 1991 Subaru Legacy

6.90 sq ft 0.641 m2 1993 Saturn Wagon

6.93 sq ft 0.644 m2 1982 Delorean DMC-12

6.94 sq ft 0.645 m2 2003 Smart Roadster

6.96 sq ft 0.647 m2 1988 Porsche 944 S

6.96 sq ft 0.647 m2 1995 Chevrolet Lumina LS

7.02 sq ft 0.652 m2 1992 BMW 325I

7.04 sq ft 0.654 m2 1991 Honda Civic EX

7.06 sq ft 0.656 m2 2004 Vauxhall VX220 Turbo

7.10 sq ft 0.660 m2 1995 Saab 900

7.11 sq ft 0.661 m2 1991 Ford Thunderbird LX

7.13 sq ft 0.662 m2 1970 Citroën SM [132]

7.14 sq ft 0.663 m2 1995 Subaru Legacy L

7.20 sq ft 0.669 m2 1995 Nissan Maxima GLE


Automobile examples of CdA[127]

CdA sqft CdA m2 Automobile model

7.34 sq ft 0.682 m2 2001 Honda Civic

7.39 sq ft 0.687 m2 1994 Honda Accord EX

7.48 sq ft 0.695 m2 1993 Chevrolet Camaro Z28

7.57 sq ft 0.703 m2 1992 Toyota Camry

7.63 sq ft 0.709 m2 1974 Citroën CX [133]

7.69 sq ft 0.714 m2 1994 Chrysler LHS

7.72 sq ft 0.717 m2 1993 Subaru Impreza

7.80 sq ft 0.725 m2 2012 Nissan Leaf SL

8.02 sq ft 0.745 m2 2005 Bugatti Veyron

8.70 sq ft 0.808 m2 1990 Volvo 740 Turbo

8.70 sq ft 0.808 m2 1992 Ford Crown Victoria

8.71 sq ft 0.809 m2 1991 Buick LeSabre Limited

8.79 sq ft 0.817 m2 1956 Citroën DS Spécial [134]

9.54 sq ft 0.886 m2 1992 Chevrolet Caprice Wagon

10.7 sq ft 0.99 m2 1992 Chevrolet Blazer

11.6 sq ft 1.08 m2 2005 Ford Escape Hybrid

11.7 sq ft 1.09 m2 1993 Jeep Grand Cherokee

12.2 sq ft 1.13 m2 1949 Nash Airflyte

16.8 sq ft 1.56 m2 2006 Hummer H3

17.4 sq ft 1.62 m2 1995 Land Rover Discovery


Automobile examples of CdA[127]

CdA sqft CdA m2 Automobile model

26.5 sq ft 2.46 m2 2003 Hummer H2

See also
 Automotive aerodynamics
 Drag (physics)
 Drag equation
 Paul Jaray

ReferencesIn physics, fluid dynamics is a subdiscipline of fluid mechanics that deals with fluid
flow—the natural science of fluids (liquids and gases) in motion. It has several subdisciplines
itself, including aerodynamics (the study of air and other gases in motion) and hydrodynamics
(the study of liquids in motion). Fluid dynamics has a wide range of applications, including
calculating forces and moments on aircraft, determining the mass flow rate of petroleum through
pipelines, predicting weather patterns, understanding nebulae in interstellar space and modelling
fission weapon detonation. Some of its principles are even used in traffic engineering, where
traffic is treated as a continuous fluid, and crowd dynamics.

Fluid dynamics offers a systematic structure—which underlies these practical disciplines—that


embraces empirical and semi-empirical laws derived from flow measurement and used to solve
practical problems. The solution to a fluid dynamics problem typically involves calculating
various properties of the fluid, such as flow velocity, pressure, density, and temperature, as
functions of space and time.

Before the twentieth century, hydrodynamics was synonymous with fluid dynamics. This is still
reflected in names of some fluid dynamics topics, like magnetohydrodynamics and
hydrodynamic stability, both of which can also be applied to gases.[1]

Contents
 1 Equations of fluid dynamics
o 1.1 Conservation laws
o 1.2 Compressible vs incompressible flow
o 1.3 Inviscid vs Newtonian and non-Newtonian fluids
o 1.4 Steady vs unsteady flow
o 1.5 Laminar vs turbulent flow
o 1.6 Subsonic vs transonic, supersonic and hypersonic flows
o 1.7 Magnetohydrodynamics
o 1.8 Other approximations
 2 Terminology in fluid dynamics
o 2.1 Terminology in incompressible fluid dynamics
o 2.2 Terminology in compressible fluid dynamics
 3 See also
o 3.1 Fields of study
o 3.2 Mathematical equations and concepts
o 3.3 Types of fluid flow
o 3.4 Fluid properties
o 3.5 Fluid phenomena
o 3.6 Applications
o 3.7 Fluid dynamics journals
o 3.8 Miscellaneous
o 3.9 See also
 4 References
 5 Further reading
 6 External links

Equations of fluid dynamics


The foundational axioms of fluid dynamics are the conservation laws, specifically, conservation
of mass, conservation of linear momentum (also known as Newton's Second Law of Motion),
and conservation of energy (also known as First Law of Thermodynamics). These are based on
classical mechanics and are modified in quantum mechanics and general relativity. They are
expressed using the Reynolds Transport Theorem.

In addition to the above, fluids are assumed to obey the continuum assumption. Fluids are
composed of molecules that collide with one another and solid objects. However, the continuum
assumption considers fluids to be continuous, rather than discrete. Consequently, properties such
as density, pressure, temperature, and flow velocity are taken to be well-defined at
infinitesimally small points, and are assumed to vary continuously from one point to another.
The fact that the fluid is made up of discrete molecules is ignored.

For fluids which are sufficiently dense to be a continuum, do not contain ionized species, and
have flow velocities small in relation to the speed of light, the momentum equations for
Newtonian fluids are the Navier–Stokes equations, which is a non-linear set of differential
equations that describes the flow of a fluid whose stress depends linearly on flow velocity
gradients and pressure. The unsimplified equations do not have a general closed-form solution,
so they are primarily of use in Computational Fluid Dynamics. The equations can be simplified
in a number of ways, all of which make them easier to solve. Some of them allow appropriate
fluid dynamics problems to be solved in closed form.[citation needed]

In addition to the mass, momentum, and energy conservation equations, a thermodynamical


equation of state giving the pressure as a function of other thermodynamic variables for the fluid
is required to completely specify the problem. An example of this would be the perfect gas
equation of state:
where p is pressure, ρ is density, Ru is the gas constant, M is molar mass and T is temperature.

Conservation laws

Three conservation laws are used to solve fluid dynamics problems, and may be written in
integral or differential form. Mathematical formulations of these conservation laws may be
interpreted by considering the concept of a control volume. A control volume is a specified
volume in space through which air can flow in and out. Integral formulations of the conservation
laws consider the change in mass, momentum, or energy within the control volume. Differential
formulations of the conservation laws apply Stokes' theorem to yield an expression which may
be interpreted as the integral form of the law applied to an infinitesimal volume at a point within
the flow.

 Mass continuity (conservation of mass): The rate of change of fluid mass inside a control volume
must be equal to the net rate of fluid flow into the volume. Physically, this statement requires
that mass is neither created nor destroyed in the control volume,[2] and can be translated into
the integral form of the continuity equation:

Above, is the fluid density, u is the flow velocity vector, and t is time. The left-hand side of the
above expression contains a triple integral over the control volume, whereas the right-hand side
contains a surface integral over the surface of the control volume. The differential form of the
continuity equation is, by the divergence theorem:

 Conservation of momentum: This equation applies Newton's second law of motion to the
control volume, requiring that any change in momentum of the air within a control volume be
due to the net flow of air into the volume and the action of external forces on the air within the
volume. In the integral formulation of this equation, body forces here are represented by fbody,

the body force per unit mass. Surface forces, such as viscous forces, are represented by ,
the net force due to stresses on the control volume surface.

The differential form of the momentum conservation equation is as follows. Here, both surface
and body forces are accounted for in one total force, F. For example, F may be expanded into an
expression for the frictional and gravitational forces acting on an internal flow.
In aerodynamics, air is assumed to be a Newtonian fluid, which posits a linear relationship
between the shear stress (due to internal friction forces) and the rate of strain of the fluid. The
equation above is a vector equation: in a three-dimensional flow, it can be expressed as three
scalar equations. The conservation of momentum equations for the compressible, viscous flow
case are called the Navier–Stokes equations.[citation needed]

 Conservation of energy: Although energy can be converted from one form to another, the total
energy in a given closed system remains constant.

Above, h is enthalpy, k is the thermal conductivity of the fluid, T is temperature, and is the
viscous dissipation function. The viscous dissipation function governs the rate at which
mechanical energy of the flow is converted to heat. The second law of thermodynamics requires
that the dissipation term is always positive: viscosity cannot create energy within the control
volume.[3] The expression on the left side is a material derivative.

Compressible vs incompressible flow

All fluids are compressible to some extent, that is, changes in pressure or temperature will result
in changes in density. However, in many situations the changes in pressure and temperature are
sufficiently small that the changes in density are negligible. In this case the flow can be modelled
as an incompressible flow. Otherwise the more general compressible flow equations must be
used.

Mathematically, incompressibility is expressed by saying that the density ρ of a fluid parcel does
not change as it moves in the flow field, i.e.,

where D/Dt is the substantial derivative, which is the sum of local and convective derivatives.
This additional constraint simplifies the governing equations, especially in the case when the
fluid has a uniform density.

For flow of gases, to determine whether to use compressible or incompressible fluid dynamics,
the Mach number of the flow is to be evaluated. As a rough guide, compressible effects can be
ignored at Mach numbers below approximately 0.3. For liquids, whether the incompressible
assumption is valid depends on the fluid properties (specifically the critical pressure and
temperature of the fluid) and the flow conditions (how close to the critical pressure the actual
flow pressure becomes). Acoustic problems always require allowing compressibility, since sound
waves are compression waves involving changes in pressure and density of the medium through
which they propagate.

Inviscid vs Newtonian and non-Newtonian fluids

Potential flow around a wing

Viscous problems are those in which fluid friction has significant effects on the fluid motion.

The Reynolds number, which is a ratio between inertial and viscous forces, can be used to
evaluate whether viscous or inviscid equations are appropriate to the problem.

Stokes flow is flow at very low Reynolds numbers, Re<<1, such that inertial forces can be
neglected compared to viscous forces.

On the contrary, high Reynolds numbers indicate that the inertial forces are more significant than
the viscous (friction) forces. Therefore, we may assume the flow to be an inviscid flow, an
approximation in which we neglect viscosity completely, compared to inertial terms.

This idea can work fairly well when the Reynolds number is high. However, certain problems
such as those involving solid boundaries, may require that the viscosity be included. Viscosity
often cannot be neglected near solid boundaries because the no-slip condition can generate a thin
region of large strain rate (known as Boundary layer) which enhances the effect of even a small
amount of viscosity, and thus generating vorticity. Therefore, to calculate net forces on bodies
(such as wings) we should use viscous flow equations. As illustrated by d'Alembert's paradox, a
body in an inviscid fluid will experience no drag force. The standard equations of inviscid flow
are the Euler equations. Another often used model, especially in computational fluid dynamics, is
to use the Euler equations away from the body and the boundary layer equations, which
incorporates viscosity, in a region close to the body.

The Euler equations can be integrated along a streamline to get Bernoulli's equation. When the
flow is everywhere irrotational and inviscid, Bernoulli's equation can be used throughout the
flow field. Such flows are called potential flows.

Sir Isaac Newton showed how stress and the rate of strain are very close to linearly related for
many familiar fluids, such as water and air. These Newtonian fluids are modelled by a constant
viscosity, depending only on the specific fluid.
However, some of the other materials, such as emulsions and slurries and some visco-elastic
materials (e.g. blood, some polymers), have more complicated non-Newtonian stress-strain
behaviours. These materials include sticky liquids such as latex, honey, and lubricants which are
studied in the sub-discipline of rheology.

Steady vs unsteady flow

Hydrodynamics simulation of the Rayleigh–Taylor instability [4]

When all the time derivatives of a flow field vanish, the flow is considered to be a steady flow.
Steady-state flow refers to the condition where the fluid properties at a point in the system do not
change over time. Otherwise, flow is called unsteady (also called transient[5]). Whether a
particular flow is steady or unsteady, can depend on the chosen frame of reference. For instance,
laminar flow over a sphere is steady in the frame of reference that is stationary with respect to
the sphere. In a frame of reference that is stationary with respect to a background flow, the flow
is unsteady.

Turbulent flows are unsteady by definition. A turbulent flow can, however, be statistically
stationary. According to Pope:[6]

The random field U(x,t) is statistically stationary if all statistics are invariant under a shift in
time.

This roughly means that all statistical properties are constant in time. Often, the mean field is the
object of interest, and this is constant too in a statistically stationary flow.

Steady flows are often more tractable than otherwise similar unsteady flows. The governing
equations of a steady problem have one dimension fewer (time) than the governing equations of
the same problem without taking advantage of the steadiness of the flow field.
Laminar vs turbulent flow

Turbulence is flow characterized by recirculation, eddies, and apparent randomness. Flow in


which turbulence is not exhibited is called laminar. It should be noted, however, that the
presence of eddies or recirculation alone does not necessarily indicate turbulent flow—these
phenomena may be present in laminar flow as well. Mathematically, turbulent flow is often
represented via a Reynolds decomposition, in which the flow is broken down into the sum of an
average component and a perturbation component.

It is believed that turbulent flows can be described well through the use of the Navier–Stokes
equations. Direct numerical simulation (DNS), based on the Navier–Stokes equations, makes it
possible to simulate turbulent flows at moderate Reynolds numbers. Restrictions depend on the
power of the computer used and the efficiency of the solution algorithm. The results of DNS
have been found to agree well with experimental data for some flows.[7]

Most flows of interest have Reynolds numbers much too high for DNS to be a viable option,[8]
given the state of computational power for the next few decades. Any flight vehicle large enough
to carry a human (L > 3 m), moving faster than 72 km/h (20 m/s) is well beyond the limit of
DNS simulation (Re = 4 million). Transport aircraft wings (such as on an Airbus A300 or Boeing
747) have Reynolds numbers of 40 million (based on the wing chord). In order to solve these
real-life flow problems, turbulence models will be a necessity for the foreseeable future.
Reynolds-averaged Navier–Stokes equations (RANS) combined with turbulence modelling
provides a model of the effects of the turbulent flow. Such a modelling mainly provides the
additional momentum transfer by the Reynolds stresses, although the turbulence also enhances
the heat and mass transfer. Another promising methodology is large eddy simulation (LES),
especially in the guise of detached eddy simulation (DES)—which is a combination of RANS
turbulence modelling and large eddy simulation.

Subsonic vs transonic, supersonic and hypersonic flows

While many terrestrial flows (e.g. flow of water through a pipe) occur at low mach numbers,
many flows of practical interest (e.g. in aerodynamics) occur at high fractions of the Mach
Number M=1 or in excess of it (supersonic flows). New phenomena occur at these Mach number
regimes (e.g. shock waves for supersonic flow, transonic instability in a regime of flows with M
nearly equal to 1, non-equilibrium chemical behaviour due to ionization in hypersonic flows) and
it is necessary to treat each of these flow regimes separately.

Magnetohydrodynamics
Main article: Magnetohydrodynamics

Magnetohydrodynamics is the multi-disciplinary study of the flow of electrically conducting


fluids in electromagnetic fields. Examples of such fluids include plasmas, liquid metals, and salt
water. The fluid flow equations are solved simultaneously with Maxwell's equations of
electromagnetism.
Other approximations

There are a large number of other possible approximations to fluid dynamic problems. Some of
the more commonly used are listed below.

 The Boussinesq approximation neglects variations in density except to calculate buoyancy


forces. It is often used in free convection problems where density changes are small.
 Lubrication theory and Hele–Shaw flow exploits the large aspect ratio of the domain to show
that certain terms in the equations are small and so can be neglected.
 Slender-body theory is a methodology used in Stokes flow problems to estimate the force on, or
flow field around, a long slender object in a viscous fluid.
 The shallow-water equations can be used to describe a layer of relatively inviscid fluid with a
free surface, in which surface gradients are small.
 The Boussinesq equations are applicable to surface waves on thicker layers of fluid and with
steeper surface slopes.
 Darcy's law is used for flow in porous media, and works with variables averaged over several
pore-widths.
 In rotating systems, the Quasi-geostrophic equations assume an almost perfect balance
between pressure gradients and the Coriolis force. It is useful in the study of atmospheric
dynamics.

Terminology in fluid dynamics


The concept of pressure is central to the study of both fluid statics and fluid dynamics. A
pressure can be identified for every point in a body of fluid, regardless of whether the fluid is in
motion or not. Pressure can be measured using an aneroid, Bourdon tube, mercury column, or
various other methods.

Some of the terminology that is necessary in the study of fluid dynamics is not found in other
similar areas of study. In particular, some of the terminology used in fluid dynamics is not used
in fluid statics.

Terminology in incompressible fluid dynamics

The concepts of total pressure and dynamic pressure arise from Bernoulli's equation and are
significant in the study of all fluid flows. (These two pressures are not pressures in the usual
sense—they cannot be measured using an aneroid, Bourdon tube or mercury column.) To avoid
potential ambiguity when referring to pressure in fluid dynamics, many authors use the term
static pressure to distinguish it from total pressure and dynamic pressure. Static pressure is
identical to pressure and can be identified for every point in a fluid flow field.

In Aerodynamics, L.J. Clancy writes:[9] To distinguish it from the total and dynamic pressures,
the actual pressure of the fluid, which is associated not with its motion but with its state, is often
referred to as the static pressure, but where the term pressure alone is used it refers to this static
pressure.
A point in a fluid flow where the flow has come to rest (i.e. speed is equal to zero adjacent to
some solid body immersed in the fluid flow) is of special significance. It is of such importance
that it is given a special name—a stagnation point. The static pressure at the stagnation point is
of special significance and is given its own name—stagnation pressure. In incompressible flows,
the stagnation pressure at a stagnation point is equal to the total pressure throughout the flow
field.

Terminology in compressible fluid dynamics

In a compressible fluid, such as air, the temperature and density are essential when determining
the state of the fluid. In addition to the concept of total pressure (also known as stagnation
pressure), the concepts of total (or stagnation) temperature and total (or stagnation) density are
also essential in any study of compressible fluid flows. To avoid potential ambiguity when
referring to temperature and density, many authors use the terms static temperature and static
density. Static temperature is identical to temperature; and static density is identical to density;
and both can be identified for every point in a fluid flow field.

The temperature and density at a stagnation point are called stagnation temperature and
stagnation density.

A similar approach is also taken with the thermodynamic properties of compressible fluids.
Many authors use the terms total (or stagnation) enthalpy and total (or stagnation) entropy. The
terms static enthalpy and static entropy appear to be less common, but where they are used they
mean nothing more than enthalpy and entropy respectively, and the prefix "static" is being used
to avoid ambiguity with their 'total' or 'stagnation' counterparts. Because the 'total' flow
conditions are defined by isentropically bringing the fluid to rest, the total (or stagnation) entropy
is by definition always equal to the "static" entropy

Drag coefficient
From Wikipedia, the free encyclopedia
Drag coefficients in fluids with Reynolds number approximately 104

In fluid dynamics, the drag coefficient (commonly denoted as: cd, cx or cw) is a dimensionless
quantity that is used to quantify the drag or resistance of an object in a fluid environment, such as
air or water. It is used in the drag equation, where a lower drag coefficient indicates the object
will have less aerodynamic or hydrodynamic drag. The drag coefficient is always associated with
a particular surface area.[1]

The drag coefficient of any object comprises the effects of the two basic contributors to fluid
dynamic drag: skin friction and form drag. The drag coefficient of a lifting airfoil or hydrofoil
also includes the effects of lift-induced drag.[2][3] The drag coefficient of a complete structure
such as an aircraft also includes the effects of interference drag.[4][5]

Contents
 1 Definition
 2 Background
o 2.1 General
 3 Aircraft
 4 Bluff and streamlined body flows
o 4.1 Concept
 4.1.1 Practical example
 5 See also
 6 Notes
 7 References

Definition
The drag coefficient is defined as:

where:

is the drag force, which is by definition the force component in the direction of the flow
velocity,[6]

is the mass density of the fluid,[7]

is the speed of the object relative to the fluid,

is the reference area.

The reference area depends on what type of drag coefficient is being measured. For automobiles
and many other objects, the reference area is the projected frontal area of the vehicle. This may
not necessarily be the cross sectional area of the vehicle, depending on where the cross section is
taken. For example, for a sphere (note this is not the surface area = ).

For airfoils, the reference area is the planform area. Since this tends to be a rather large area
compared to the projected frontal area, the resulting drag coefficients tend to be low: much lower
than for a car with the same drag and frontal area, and at the same speed.

Airships and some bodies of revolution use the volumetric drag coefficient, in which the
reference area is the square of the cube root of the airship volume (volume to the two-thirds
power). Submerged streamlined bodies use the wetted surface area.

Two objects having the same reference area moving at the same speed through a fluid will
experience a drag force proportional to their respective drag coefficients. Coefficients for
unstreamlined objects can be 1 or more, for streamlined objects much less.

Background
Flow around a plate, showing stagnation.

Main article: Drag equation

The drag equation:

is essentially a statement that the drag force on any object is proportional to the density of the
fluid and proportional to the square of the relative speed between the object and the fluid.

Cd is not a constant but varies as a function of speed, flow direction, object position, object size,
fluid density and fluid viscosity. Speed, kinematic viscosity and a characteristic length scale of
the object are incorporated into a dimensionless quantity called the Reynolds number or . is
thus a function of . In compressible flow, the speed of sound is relevant and is also a
function of Mach number .

For a certain body shape, the drag coefficient only depends on the Reynolds number , Mach
number and the direction of the flow. For low Mach number , the drag coefficient is
independent of Mach number. Also, the variation with Reynolds number within a practical
range of interest is usually small, while for cars at highway speed and aircraft at cruising speed
the incoming flow direction is also more-or-less the same. So the drag coefficient can often be
treated as a constant.[8]

For a streamlined body to achieve a low drag coefficient, the boundary layer around the body
must remain attached to the surface of the body for as long as possible, causing the wake to be
narrow. A high form drag results in a broad wake. The boundary layer will transition from
laminar to turbulent providing the Reynolds number of the flow around the body is high enough.
Larger velocities, larger objects, and lower viscosities contribute to larger Reynolds numbers.[9]
Drag coefficient Cd for a sphere as a function of Reynolds number Re, as obtained from laboratory
experiments. The solid line is for a sphere with a smooth surface, while the dashed line is for the case of
a rough surface. The numbers along the line indicate several flow regimes and associated changes in the
drag coefficient:
•2: attached flow (Stokes flow) and steady separated flow,
•3: separated unsteady flow, having a laminar flow boundary layer upstream of the separation, and
producing a vortex street,
•4: separated unsteady flow with a laminar boundary layer at the upstream side, before flow
separation, with downstream of the sphere a chaotic turbulent wake,
•5: post-critical separated flow, with a turbulent boundary layer.

For other objects, such as small particles, one can no longer consider that the drag coefficient
is constant, but certainly is a function of Reynolds number.[10][11][12] At a low Reynolds number,
the flow around the object does not transition to turbulent but remains laminar, even up to the
point at which it separates from the surface of the object. At very low Reynolds numbers,
without flow separation, the drag force is proportional to instead of ; for a sphere this is
known as Stokes law. Reynolds number will be low for small objects, low velocities, and high
viscosity fluids.[9]

A equal to 1 would be obtained in a case where all of the fluid approaching the object is
brought to rest, building up stagnation pressure over the whole front surface. The top figure
shows a flat plate with the fluid coming from the right and stopping at the plate. The graph to the
left of it shows equal pressure across the surface. In a real flat plate, the fluid must turn around
the sides, and full stagnation pressure is found only at the center, dropping off toward the edges
as in the lower figure and graph. Only considering the front side, the of a real flat plate would
be less than 1; except that there will be suction on the back side: a negative pressure (relative to
ambient). The overall of a real square flat plate perpendicular to the flow is often given as
1.17.[citation needed] Flow patterns and therefore for some shapes can change with the Reyd
examples==

General

In general, is not an absolute constant for a given body shape. It varies with the speed of
airflow (or more generally with Reynolds number ). A smooth sphere, for example, has a
that varies from low values for laminar flow to 0.47 for turbulent flow.
Shapes

cd Item

0.001 laminar flat plate parallel to the flow ( )

0.005 turbulent flat plate parallel to the flow ( )

0.075 Pac-car

0.1 smooth sphere ( )

0.47 smooth sphere ( )

0.15 Schlörwagen 1939 [13]

0.186-0.189 Volkswagen XL1 2014

0.19 General Motors EV1 1996[14]

0.24 Tesla Model S[15]

0.25 Toyota Prius (3rd Generation)

0.26 BMW i8

0.26 Nissan GT-R (2011-2014)

0.27 Nissan GT-R (2007-2010)

0.28 Mercedes-Benz CLA-Class Type C 117.[16]

0.29 Mazda3 (2007) [17]

0.295 bullet (not ogive, at subsonic velocity)

0.3 Saab 92 (1949), Audi 100 C3 (1982)

0.324 Ford Focus Mk2/2.5 (2004-2011, Europe)

0.48 rough sphere ( ),


Volkswagen Beetle[18][19]
Shapes

cd Item

0.58 Jeep Wrangler TJ (1997-2005)[20]

0.75 a typical model rocket[21]

1.0 coffee filter, face-up[22]

1.0 road bicycle plus cyclist, touring position[23]

1.0–1.1 skier

1.0–1.3 wires and cables

1.0–1.3 man (upright position)

1.1-1.3 ski jumper[24]

1.28 flat plate perpendicular to flow (3D) [25]

1.3–1.5 Empire State Building

1.8–2.0 Eiffel Tower

1.98–2.05 flat plate perpendicular to flow (2D)

2.1 a smooth brick[citation needed]

Aircraft
As noted above, aircraft use their wing area as the reference area when computing , while
automobiles (and many other objects) use frontal cross sectional area; thus, coefficients are not
directly comparable between these classes of vehicles. In the aerospace industry, the drag
coefficient is sometimes expressed in drag counts where 1 drag count = 0.0001 of a .[26]

Aircraft[27]

cd Aircraft type

0.021 F-4 Phantom II


(subsonic)

0.022 Learjet 24

0.024 Boeing 787[28]

0.0265 Airbus A380[29]

0.027 Cessna 172/182

0.027 Cessna 310

0.031 Boeing 747

F-4 Phantom II
0.044
(supersonic)

0.048 F-104 Starfighter

0.095 X-15 (Not confirmed)

Bluff and streamlined body flows


Concept

Drag, in the context of fluid dynamics, refers to forces that act on a solid object in the direction
of the relative fluid flow velocity. The aerodynamic forces on a body come primarily from
differences in pressure and viscous shearing stresses. Thereby, the drag force on a body could be
divided into two components, namely frictional drag (viscous drag) and pressure drag (form
drag). The net drag force could be decomposed as follows:

Flow across an airfoil showing the relative impact of drag force to the direction of motion of fluid over
the body. This drag force gets divided into frictional drag and pressure drag. The same airfoil is
considered as a streamlined body if friction drag (viscous drag) dominates pressure drag and is
considered a bluff body when pressure drag (form drag) dominates friction drag.
where:

is the pressure drag coefficient,

is the friction drag coefficient,

= Tangential direction to the surface with area dA,

= Normal direction to the surface with area dA,

is the shear Stress acting on the surface dA,

is the pressure far away from the surface dA,

is pressure at surface dA,

is the unit vector in direction normal to the surface dA, forming a unit vector

Therefore, when the drag is dominated by a frictional component, the body is called a
streamlined body; whereas in the case of dominant pressure drag, the body is called a bluff body.
Thus, the shape of the body and the angle of attack determine the type of drag. For example, an
airfoil is considered as a body with a small angle of attack by the fluid flowing across it. This
means that it has attached boundary layers, which produce much less pressure drag.

Trade-off relationship between pressure drag and friction drag

The wake produced is very small and drag is dominated by the friction component. Therefore,
such a body (here an airfoil) is described as streamlined, whereas for bodies with fluid flow at
high angles of attack, boundary layer separation takes place. This mainly occurs due to adverse
pressure gradients at the top and rear parts of an airfoil.
Due to this, wake formation takes place, which consequently leads to eddy formation and
pressure loss due to pressure drag. In such situations, the airfoil is stalled and has higher pressure
drag than friction drag. In this case, the body is described as a bluff body.

A streamlined body looks like a fish (Tuna, Oropesa, etc.) or an airfoil with small angle of
attack, whereas a bluff body looks like a brick, a cylinder or an airfoil with high angle of attack.
For a given frontal area and velocity, a streamlined body will have lower resistance than a bluff
body. Cylinders and spheres are taken as bluff bodies because the drag is dominated by the
pressure component in the wake region at high Reynolds number.

To reduce this drag, either the flow separation could be reduced or the surface area in contact
with the fluid could be reduced (to reduce friction drag). This reduction is necessary in devices
like cars, bicycle, etc. to avoid vibration and noise production.

Practical example

Aerodynamic design of cars has evolved from 1920s to the end of 20th century. This change in
design from a bluff body to a more streamlined body reduced the drag coefficient from about
0.95 to 0.30.

Time history of Aerodynamic drag of cars in comparison with change in geometry of streamlined bodies
(bluff to streamline).
See also
 Automotive aerodynamics
 Automobile drag coefficient
 Ballistic coefficient
 Drag crisis
 Zero-lift drag coefficient

Notes
1.

 McCormick, Barnes W. (1979): Aerodynamics, Aeronautics, and Flight Mechanics. p. 24, John Wiley &
Sons, Inc., New York, ISBN 0-471-03032-5
  Clancy, L. J.: Aerodynamics. Section 5.18

  Abbott, Ira H., and Von Doenhoff, Albert E.: Theory of Wing Sections. Sections 1.2 and 1.3

  "NASA’s Modern Drag Equation". Wright.nasa.gov. 2010-03-25. Retrieved 2010-12-07.

  Clancy, L. J.: Aerodynamics. Section 11.17

  See lift force and vortex induced vibration for a possible force components transverse to the flow
direction.

  Note that for the Earth's atmosphere, the air density can be found using the barometric formula. Air
is 1.293 kg/m3 at 0 °C and 1 atmosphere.

  Clancy, L. J.: Aerodynamics. Sections 4.15 and 5.4

  Clancy, L. J.: Aerodynamics. Section 4.17

 Clift R., Grace J. R., Weber M. E.: Bubbles, drops, and


particles. Academic

Anda mungkin juga menyukai