You are on page 1of 30

Governing Equations for Fundamental Aerodynamics

John D. Anderson Jr.


Aeronautics Division, National Air and Space Museum, Smithsonian Institution, Washington, DC, USA

continuous substance. The molecules impact the body surface


1 Fundamental Principles 1 so frequently that the body cannot distinguish the individual
2 Potential Flow Equations 8 molecular collisions, and the surface feels the fluid as a con-
tinuous medium. Such flow is called continuum flow. The
3 Euler Equations 15
other extreme is where λ is on the same order as the body
4 Navier–Stokes Equations 18 scale; here, the gas molecules are spaced so far apart (rel-
5 Boundary-Layer Equations 21 ative to d) that collisions with the body surface occur only
6 Chemically Reacting Flow Equations 25 infrequently, and the body surface can feel distinctly each
References 29 molecular impact. Such flow is called free molecular flow. For
manned flight, vehicles such as the space shuttle encounter
Further Reading 29
free molecular flow at the extreme outer edge of the atmo-
sphere, where the air density is so low that λ becomes on
the order of the shuttle size. There are intermediate cases,
where flows can exhibit some characteristics of both con-
1 FUNDAMENTAL PRINCIPLES tinuum and free molecular flows; such flows are generally
labeled “low-density flows” in contrast to continuum flow.
1.1 Inviscid and viscous flows: the distinction By far, however, the vast majority of practical aerodynamic
applications involve continuum flows. Low-density and free
For the purposes of analysis, discussion, and understanding,
molecular flows are just a small part of the total spectrum of
we divide and classify the real natural world of aerodynamics
aerodynamics.
into various intellectually defined compartments and cate-
For the purpose of intellectual discussion, we divide con-
gories. Underlying all of these is the idea that usually (but
tinuum flow into two categories: viscous flow and inviscid
not always) the flow medium is a continuum. To understand
flow. To define these categories, note that in a liquid or a gas,
this concept, consider the flow over a body, say, for example,
the molecules move about in a very random fashion, and in so
a circular cylinder of diameter, d. Also, consider the fluid to
doing they transport their mass, momentum, and energy from
consist of individual molecules that are in random motion.
one location to another in the fluid. This transport on a molec-
The mean distance that a molecule travels between collisions
ular scale gives rise to the phenomena of mass diffusion,
with neighboring molecules is defined as the mean-free path
viscosity (friction), and thermal conduction. All real flows
λ. If λ is orders of magnitude smaller than the scale of the
exhibit the effects of these transport phenomena; such flows
body measured by d, then the flow appears to the body as a
are called viscous flows. In contrast, a flow that is assumed
to involve no friction, thermal conduction, or diffusion is
called an inviscid flow. Inviscid flows do not truly exist in
nature; however, there are many practical aerodynamic flows
(more than one would think) where the influence of transport

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
2 Fundamentals of Fluid Flows

phenomena is small, and we can model the flow as being is locally:


inviscid.
Theoretically, inviscid flow is approached in the limit Subsonic if M < 1
as the Reynolds number goes to infinity. Reynolds number Sonic if M = 1
reflects the ratio of inertia to friction forces in a moving fluid, Supersonic if M > 1
which is given by
Looking at the whole field simultaneously, four different
Re = ρV/µ (1) speed regimes can be identified using Mach number as the
criterion:
where ρ is the fluid density, V is the flow velocity,  is a char-
acteristic length, and µ is the viscosity coefficient. However, 1. Subsonic flow (M < 1 everywhere). A flow field is defined
for practical problems, many flows with high but finite Re can as subsonic if the Mach number is less than 1 at
be assumed to be inviscid. For such flows, the influence of every point. Subsonic flows are characterized by smooth
friction, thermal conduction, and diffusion is limited to a very streamlines (no discontinuity in slope), as shown in Fig-
thin region adjacent to the body surface called the boundary ure 1a. Moreover, since the flow velocity is everywhere
layer (Section 5), and the remainder of the flow outside this less than the speed of sound, disturbances in the flow
thin region is essentially inviscid. For flows over slender bod- (say, the sudden deflection of the trailing edge of the
ies, such as an airfoil, inviscid theory adequately predicts the airfoil in Figure 1a) propagate both upstream and down-
pressure distribution and lift on the body and gives a valid stream, and are felt throughout the entire flow field. A
representation of the streamlines and flow field away from freestream Mach number M∞ less than 1 does not guar-
the body. However, because friction (shear stress) is a major antee a totally subsonic flow over the body. In expanding
source of aerodynamic drag, inviscid theories by themselves over an aerodynamic shape, the flow velocity increases
cannot adequately predict total drag. above the freestream value, and if M∞ is close enough
to 1, the local Mach number may become supersonic
in certain regions of the flow. This gives rise to a rule of
1.2 Incompressible and compressible flows: thumb that M∞ < 0.8 for subsonic flow over slender bod-
the distinction ies. For blunt bodies, M∞ must be even lower to ensure
totally subsonic flow.
A flow in which the density ρ is constant is called incom- 2. Transonic flow (mixed regions where M < 1 and M > 1).
pressible. In contrast, a flow where the density is variable is As stated above, if M∞ is subsonic but is near unity,
called compressible. All flows, to a greater or lesser extent, the flow can become locally supersonic (M > 1). This
are compressible; truly incompressible flow, where the den- is shown in Figure 1b, which shows pockets of super-
sity is precisely constant, does not occur in nature. However, sonic flow over both the top and the bottom surfaces
analogous to our discussion of inviscid flow, there are a of the airfoil, terminated by weak shock waves behind
number of aerodynamic problems that can be modeled as which the flow becomes subsonic again. Moreover, if
being incompressible without any detrimental loss of accu- M∞ is increased slightly above unity, a bow shock wave
racy. For example, the flow of homogeneous liquids is treated is formed in front of the body; behind this shock wave,
as incompressible, and hence most problems involving hydro- the flow is locally subsonic, as shown in Figure 1c. This
dynamics assume ρ = constant. Also, the flow of gases at a subsonic flow subsequently expands to a low supersonic
low Mach number is essentially incompressible; for M < 0.3, value over the airfoil. Weak shock waves are usually
it is always safe to assume ρ = constant. This was the flight generated at the trailing edge, sometimes in a “fishtail”
regime of all airplanes from the Wright brothers’ first flight in pattern as shown in Figure 1c. The flow fields shown in
1903 to just prior to World War II. It is still the flight regime Figure 1b and c are characterized by mixed subsonic–
of most small, general aviation aircraft of today. Hence, there supersonic flows and are dominated by the physics of
exists a large bulk of aerodynamic experimental and theoret- both types of flow. Hence, such flow fields are called tran-
ical data for incompressible flows. sonic flows. Again, as a rule of thumb for slender bodies,
Of all the ways of subdividing and describing different transonic flows occur for freestream Mach numbers in
aerodynamic flows, the distinction based on the Mach number the range 0.8 < M∞ < 1.2.
is probably the most prevalent. If M is the local Mach number 3. Supersonic flow (M > 1 everywhere). A flow field is
at an arbitrary point in a flow field, then by definition the flow defined as supersonic if the Mach number is greater than

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 3

M ∞ < 0.8

(a)

M >1

0.8 <M∞ < 1


M >1

(b)

k
o ck

oc
sh
Bow sh

ge
ed
g-
M <1 M >1 ilin
Tra

1 <M∞ < 1.2

(c)
ve

1
M>
wa

Expansion wave
k
oc
Sh

M∞ > 1.2 θ

(d)

Thin, hot
shock layer
M∞ > 5 with viscous
interaction
and chemical
(e) reactions

Figure 1. Different regimes of flow. (a) Subsonic flow; (b) transonic flow with M∞ < 1; (c) transonic flow with M∞ > 1; (d) Supersonic
flow; and (e) hypersonic flow.

1 at every point. Supersonic flows are frequently charac- flow over a sharp-nosed wedge; the flow remains super-
terized by the presence of shock waves across which the sonic behind the oblique shock wave from the tip. Also
flow properties and streamlines change discontinuously shown are the distinct expansion waves, which are com-
(in contrast to the smooth, continuous variations in sub- mon in supersonic flow. Again, the listing of M∞ > 1.2
sonic flows). This is shown in Figure 1d for supersonic is strictly a rule of thumb. For example, in Figure 1d,

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
4 Fundamentals of Fluid Flows

if θ is made large enough, the oblique shock wave will which we can apply the fundamental principles stated in
detach from the tip of the wedge and will form a strong, item 1.
curved bow shock ahead of the wedge with a substan- 3. Apply the fundamental physical principles listed in item
tial region of subsonic flow behind the wave. Hence, the 1 to the model of the fluid determined in item 2 in order to
totally supersonic flow shown in Figure 1d is destroyed obtain mathematical equations that properly describe the
if θ is too large for a given M∞ . This shock detachment physics of the flow. In turn, use these fundamental equa-
phenomenon can occur at any value of M∞ > 1, but the tions to analyze any particular aerodynamic flow problem
value of θ at which it occurs increases as M∞ increases. of interest.
In a supersonic flow, because the local flow velocity is
greater than the speed of sound, disturbances created at Three different models of the flow have been used suc-
some point in the flow cannot work their way upstream cessfully throughout the modern evolution of aerodynamics.
(in contrast to subsonic flow). This property is one of They are (1) (2) finite control volume, (3) infinitesimal fluid
the most significant physical differences between sub- element, and (4) molecular.
sonic and supersonic flows. This is the basic reason why Let us imagine a closed volume within a finite region of the
shock waves occur in supersonic flows, but do not occur flow. This volume defines a control volume V and a control
in steady subsonic flows. surface S; control surface is the closed surface that bounds the
4. Hypersonic flow (very high supersonic speeds). Refering control volume. The control volume may be fixed in space
to the wedge shown in Figure 1d, assume θ is a given, with the fluid moving through it. Alternatively, the control
fixed value. As M∞ increases above 1, the strength of volume may be moving with the fluid such that the same
the shock wave increases, leading to higher tempera- fluid particles are always inside it. In either case, the control
tures in the region between the shock and the body (the volume is a reasonably large, finite region of the flow. The
shock layer). If M∞ is sufficiently large, the shock layer fundamental physical principles are applied to the fluid inside
becomes very thin, and interactions between the shock the control volume and to the fluid crossing the control surface
wave and the viscous boundary layer on the surface occur. (if the control volume is fixed in space). Therefore, instead
Also, the shock layer temperature becomes high enough of looking at the whole flow field at once, with the control
that chemical reactions occur in the air. The O2 and N2 volume model we limit our attention to just the fluid in the
molecules are torn apart, that is, the gas molecules disso- finite region of the volume itself.
ciate. When M∞ becomes large enough such that shock Alternatively, let us imagine an infinitesimally small fluid
and viscous layer interactions, and/or chemically react- element in the flow, with a differential volume dV. The fluid
ing effects begin to dominate the flow (Figure 1e), the element is infinitesimal in the same sense as differential cal-
flow field is called hypersonic. A somewhat arbitrary culus; however, it is large enough to contain a huge number of
but frequently used rule of thumb for hypersonic flow molecules so that it can be viewed as a continuous medium.
is M∞ > 5. The fluid element may be fixed in space with the fluid moving
through it. Alternatively, it may be moving along a stream-
1.3 Control volumes and fluid elements line with velocity V equal to the flow velocity at each point.
Again, instead of looking at the whole flow field at once, the
The fundamental equations of aerodynamics are obtained fundamental physical principles are applied to just the fluid
from the following philosophical approach: element itself.
A third model of the flow can be a microscopic approach
1. Invoke the following three fundamental physical prin- that pertains to the fact that the motion of a fluid is a ramifica-
ciples that are deeply entrenched in our macroscopic tion of the mean motion of its atoms and molecules. Here, the
observations of nature: fundamental laws of nature are applied directly to the atoms
a. Mass is conserved (i.e., mass can be neither created and molecules, using suitable statistical averaging to define
nor destroyed). the resulting fluid properties. This approach is in the purview
b. Newton’s second law: force = mass × acceleration. of kinetic theory, which is a very elegant method with many
c. Energy is conserved – it can only change from one advantages in the long run.
form to another form.
2. Determine a suitable model for the fluid. A fluid is a 1.4 Continuity equation
squishy substance, and therefore it is usually more dif-
ficult to describe than a well-defined solid body. Hence, The basic equations of aerodynamics are statements of the
we have to adopt a reasonable model of the fluid to three fundamental principles listed in Section 1.3. These

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 5

principles are fundamental in nature; they are stand-alone right-hand side represent the force exerted on the fluid inside
principles that cannot be derived from other more fundamen- the control volume as it is momentarily flowing through the
tal concepts. volume. This force is the sum of the pressure force due to
Fundamental principle: Mass is conserved (i.e., mass can the integral of the pressure distribution exerted on the control
be neither created nor destroyed). surface, the body force exerted on the fluid inside the con-
Applying this principle to the model of a finite control trol volume, and the force due to viscous stresses (friction)
volume of volume V and surface area S fixed in space, the exerted over the control surface. In equation (4), p is the pres-
resulting equation is sure, f is the body force per unit mass, and F viscous is the net
surface force due to friction. Equation (4) is the integral form
∂ ⌠⌠⌠ ⌠⌠ of the momentum equation; it is a vector equation.

⌡ ⌡ ρdV + ⌡⌡ ρV ·dS = 0 (2) Equation (4) can be expressed in terms of three scalar par-
∂t V S
tial differential equations. For instance, adopting a Cartesian
(x, y, z) coordinate system, we have
where the first integral over the control volume V is the time
rate of change of mass inside the control volume and the ∂(ρu) ∂p
second integral over the control surface S is the net flux of + ∇·(ρuV ) = − + ρfx + (Fx )viscous (5)
∂t ∂x
mass out of the control volume through the control surface.
Equation (2) is the integral form of the continuity equation. ∂(ρv) ∂p
Applying the divergence theorem from vector calculus + ∇·(ρvV ) = − + ρfy + (Fy )viscous (6)
∂t ∂y
to equation (2), we obtain the following partial differential
equation: ∂(ρw) ∂p
+ ∇·(ρwV ) = − + ρfz + (Fz )viscous (7)
∂t ∂z
∂ρ
+ ∇·(ρV ) = 0 (3)
∂t Equations (5)–(7) are respectively the x, y, and z com-
ponents of the momentum equation in partial differential
Equation (2) is a differential form of the continuity equation. equation form.
In both equations (2) and (3), ρ is the flow density and V is the
flow velocity. Both equations address the same subject – the
mass is conserved. Equation (2), however, applies over a finite 1.6 Energy equation
volume in space, and equation (3) applies in the neighborhood
of a point in the flow. Fundamental principle: Energy is conserved; it can only
change from one form to another form.
This principle is simply a statement of the first law of
1.5 Momentum equation thermodynamics to the fluid flow, namely, that the change in
energy of a system is equal to the heat added to, and the work
Fundamental principle: Newton’s second law: done on the system. Applying this principle to a finite control
force = mass × acceleration, or more fundamentally volume fixed in space, we have
that force = time rate of change of momentum.
Applying this fundamental principle to a finite control ⌠⌠⌠ ⌠⌠ ⌠⌠⌠
volume fixed in space, we have ⌡⌡⌡ q̇ρdV + Q̇viscous − ⌡⌡ pV ·dS + ⌡⌡⌡ ρ(f ·V )dV
V S V

 
∂ ⌠⌠⌠ ⌠⌠ ∂ ⌠⌠⌠ V2
ρV dV + ⌡⌡ (ρV ·dS)V + Ẇviscous = +
∂t ⌡V⌡⌡ ∂t ⌡V⌡⌡
ρ e dV
S
2
 
⌠⌠ ⌠⌠⌠ ⌠⌠ V2
= − ⌡⌡ pdS + ⌡⌡⌡ ρf dV + F viscous (4) + ⌡⌡ ρ e + V ·dS (8)
S V S
2

where the first integral on the left-hand side is the time rate Equation (8) is an integral form of the energy equation. On
of change of momentum inside the control volume and the the left side, the first term is the volumetric addition of heat
second integral is the net flux of momentum out of the con- to the fluid inside the control volume from an outside source
trol volume through the control surface. The terms on the such as radiation; the second term is the heat addition due to

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
6 Fundamentals of Fluid Flows

the thermal conduction across the control surface; the third = −∇·(pV ) + ρ(f ·V ) (13)
term is the rate of work done by pressure exerted on the con-
trol surface; the fourth term is the rate of work done by body In modern terminology, the system of equations (3) and
forces; and the fifth term is the rate of work done by viscous (10)–(13) are labeled as the Euler equations, although classi-
stresses exerted on the control surface. The first integral on cally only the system of equations (10)–(12) – the momentum
the right side of equation (8) is the time rate of change of total equation for an inviscid flow – had been called the Euler
energy inside the control volume and the second integral is equation.
the flux of total energy across the control surface.
Equation (8) can be expressed in partial differential equa-
tion form as 1.8 Pathlines, streamlines, and streaklines
      
∂ V2 V2 A pathline is the path through space traced out by a fluid
ρ e+ + ∇· ρ e + V
∂t 2 2 element moving in the flow. Suppose when you are watching a
given fluid element as it flows downstream, the path followed
= ρq̇ − ∇·(pV ) + ρ(f ·V ) + Q̇viscous + Ẇviscous

(9)
by your eyes is the pathline that is traced out by the fluid
element.
where Q̇viscous and Ẇviscous

represent the proper forms of the Imagine velocity vectors attached to all the fluid elements
viscous terms, to be expressed in Section 4. Equation (9) is as they move through the flow, and the array of these velocity
a partial differential equation, which relates the flow-field vectors distributed throughout space at a specific instant. A
variables at a given point in space. streamline is a curve whose tangent at any point is in the
direction of the velocity vector at that point.
Consider a fixed point in a flow field. Consider all the indi-
1.7 Classifications: Euler equations and vidual fluid elements that pass through that point over a given
Navier–Stokes equations time interval. Imagine all these fluid elements connected with
each other, such as a string of elephants connected trunk to
The governing equations for an unsteady, compressible, tail. The line that connects these fluid elements is a streakline.
three-dimensional viscous flow are given by equations (2)– In general, a flow field can be unsteady, where by defi-
(8). In modern terminology, these equations – the continuity, nition the flow field properties, including the magnitude and
momentum, and energy equations – are labeled as the Navier– direction of the velocity, at any given point are functions
Stokes equations, although classically only the momentum of time. For an unsteady flow, the pathline, streamline, and
equation with its viscous terms included had been called the streakline through a given point in the flow are all different
Navier–Stokes equation. curves. In contrast, a steady flow is one where the flow-field
For an inviscid flow, the terms in the above governing properties at any given point do not vary with time; they are
equations dealing with viscous effects drop out. Since the the same fixed values independent of time. For a steady flow,
continuity equation deals only with mass conservation, it pathlines, streamlines, and streaklines through any point are
remains the same. However, the momentum equation (equa- identical curves in space. For a steady flow, therefore, there
tions (5)–(7)) is now written as is no difference between these curves, and for such flows, the
term streamline is the one that is most used to give a clear
∂(ρu) ∂p picture of the flow.
+ ∇·(ρuV ) = − + ρfx (10)
∂t ∂x

∂(ρv) ∂p 1.9 Vorticity


+ ∇·(ρvV ) = − + ρfy (11)
∂t ∂y
As a moving fluid element translates along a streamline, it
∂(ρw) ∂p may also rotate and become distorted in shape. The amount
+ ∇·(ρwV ) = − + ρfz (12)
∂t ∂z of rotation and distortion depends on the velocity field. The
rotational angular velocity of the fluid element is a vector
The energy equation (9) becomes quantity given by the symbol ω. By definition, the vorticity ξ
is twice the angular velocity,
      
∂ V2 V2
ρ e+ + ∇· ρ e + V
∂t 2 2 ξ = 2ω (14)

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 7

and is equal to the curl of the velocity, three unknowns, we can deal with the velocity potential as one
unknown; therefore, the solution of only one equation for the
ξ =∇×V (15) flow field is required. This is why, in theoretical aerodynam-
ics, we make a distinction between irrotational and rotational
In a velocity field, the curl of the velocity is equal to the flows, and why the analysis of irrotational flow is simpler than
vorticity. that of rotational flows. This is also why irrotational flow is
The above leads to two important definitions: often identified as potential flow.

1. If (∇ × V ) =/ 0 at every point in a flow, the flow is called 1.11 Stream function


rotational. This implies that the fluid elements have a
finite angular velocity. The stream function, ψ, such as the velocity potential, is a
2. If (∇ × V ) = 0 at every point in a flow, the flow is called scalar function from which the flow velocity can be obtained
irrotational. This implies that the fluid elements have no by differentiation. For example, in Cartesian coordinates,
angular velocity; rather their motion through space is a
pure translation. ∂ψ ∂ψ
u= v=− (18)
∂y ∂x
Irrotational flows are much easier to analyze than rota-
tional flows. At first glance, irrotational flow may appear to Note that the derivatives are taken perpendicular to the
be so special that its applications are limited. Amazingly, direction of the velocity components, in contrast to the case
such is not the case. There are a large number of practi- with the velocity potential. Also, a stream function, except in
cal aerodynamic problems where the flow field is essentially several special cases, is defined for only a two-dimensional
irrotational, for example, the subsonic flow over airfoils, the flow, whereas the velocity potential can be defined for a three-
supersonic flow over slender bodies at small angle of attack, dimensional flow.
and the subsonic–supersonic flow through nozzles. For such Along a streamline in a two-dimensional flow, the value
cases, there is generally a thin boundary layer of viscous flow of the stream function is constant. Hence, the equations of a
immediately adjacent to the surface; in this viscous region, streamline can be expressed as ψ = constant, where the value
the flow is highly rotational. However, outside this boundary of the constant is different for different streamlines.
layer, the flow is frequently irrotational. As a result, the study A stream function can be defined for a rotational flow as
of irrotational flow is an important aspect of aerodynamics. well as an irrotational flow. This is in contrast to a velocity
potential, which can be defined only for an irrotational flow.

1.10 Velocity potential


1.12 Circulation
For an irrotational flow, there exists at each point a scalar
function φ such that the velocity at that point is given by In this section, we introduce a tool that is fundamental to
the calculation of aerodynamic lift, namely, circulation. This
V = ∇φ (16) tool was used independently by Frederick Lanchester (1878–
1946) in England, Wilheim Kutta (1867–1944) in Germany,
We denote φ as the velocity potential. If the velocity potential and Nikolai Joukowski (1847–1921) in Russia to create a
for a given flow is known, then the velocity can be obtained breakthrough in the theory of aerodynamic lift at the turn of
directly from differentiation of the velocity potential. For the twentieth century. The purpose of this section is only to
example, in Cartesian coordinates, define circulation and relate it to vorticity.
Consider a closed curve C in a flow field. Let V and ds
∂φ ∂φ ∂φ be the velocity and directed line segment, respectively, at a
u= v= w= (17)
∂x ∂y ∂z point on C. The circulation, denoted by , is defined as
Note that the flow-field velocities are obtained by differenti-
ating φ in the same direction as the velocities. ⌠
 ≡ − ⌡ V ·ds (19)
When a flow field is irrotational, hence allowing a velocity C
potential to be defined, there is a tremendous simplification.
Instead of dealing with the velocity components (say, u, v, The circulation is simply the negative of the line integral of
and w) as unknowns, hence requiring three equations for these velocity around a closed curve in the flow; it is a kinematic

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
8 Fundamentals of Fluid Flows

property depending only on the velocity field and the choice 2 POTENTIAL FLOW EQUATIONS
of the curve C. By mathematical convention, the positive
sense of the line integral is counterclockwise. However, in 2.1 Definition of potential flow
aerodynamics, it is convenient to consider a positive cir-
culation as being clockwise. Hence, a minus sign appears The velocity potential, φ, is defined in Section 1.10 as a
in the definition given by equation (19) to account for the scalar function from which the vector velocity is obtained
positive-counterclockwise sense of the integral and for the as V = ∇φ. The velocity potential can be defined for an irro-
positive-clockwise sense of circulation. tational flow only (Section 1.10). Therefore, potential flow
The use of the word circulation to label the integral in can be defined as a flow for which a velocity potential exists,
equation (19) may be somewhat misleading because it leaves hence potential flow is irrotational flow.
a general impression of something moving around in a loop. Frequently, the term potential flow is used to describe
Indeed, according to the American Heritage Dictionary of the incompressible irrotational flow. However, this is too restric-
English Language, the first definition given to the word “cir- tive because a velocity potential can also be defined for
culation” is “movement in a circle or a circuit.” However, in irrotational compressible flow. Expressing the flow velocity
aerodynamics, circulation has a very precise technical mean- in Cartesian coordinates, we have
ing, namely, equation (19). It does not necessarily mean that
the fluid elements are moving around in circles within this V = ui + vj + wk (23)
flow field – a common early misconception of new learners
of aerodynamics. Rather, when circulation exists in a flow, it where i, j, and k are unit vectors along the x, y, and z axes
simply means that the line integral in equation (19) is finite. respectively, and u, v, and w are the x, y, and z components
For example, if an airfoil is generating lift, the circulation of velocity, respectively. Since
taken around a closed curve enclosing the airfoil will be finite,
although the fluid elements are by no means executing circles ∂φ ∂φ ∂φ
V = ∇φ = i +j +k (24)
around the airfoil. ∂x ∂y ∂z
Circulation is also related to vorticity as follows.
we have, comparing equations (23) and (24),
⌠ ⌠⌠
 ≡ − ⌡ V ·ds = − ⌡⌡ (∇ × V )·dS (20) ∂φ
C S
u= (25)
∂x
Hence, the circulation about a curve C is equal to the ∂φ
vorticity integrated over any open surface bounded by C. v= (26)
∂y
This leads to the immediate result that if the flow is irro-
tational everywhere within the contour of integration (i.e., if ∂φ
∇ × V = 0 over any surface bounded by C), then  = 0. A w= (27)
∂z
related result is obtained by letting the curve C shrink to an
infinitesimal size, and denoting the circulation around this Equations (25)–(27) illustrates one of the most powerful
infinitesimally small curve by d. Then, in the limit as C advantages of defining and obtaining the velocity potential for
becomes infinitesimally small, equation (20) yields a given flow, that is, the three unknown velocities u, v, and
w are replaced by only one unknown, namely, φ. Hence, the
d = −(∇ × V )·dS = −(∇ × V )·ndS (21) theoretical analysis of potential flows is inherently simpler
than that for a nonpotential flow.
or

d 2.2 Bernoulli’s equation


(∇ × V )·n = − (22)
dS
For an incompressible flow, the integration of the Euler equa-
where dS is the infinitesimal area enclosed by the infinitesimal tions (Section 3) along a streamline yields a single algebraic
curve C. Equation (22) states that at a point P in a flow, the equation relating pressure and velocity along the streamline:
component of vorticity normal to dS is equal to the negative of
the “circulation per unit area,” where the circulation is taken 1
around the boundary of dS. p + ρV2 = constant (28)
2

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 9

Equation (28) is called Bernoulli’s equation; it reflects the Equation (31) is Laplace’s equation – one of the most
Bernoulli principle that, along a given streamline, as the pres- famous and extensively studied equations in mathematical
sure decreases, the velocity increases, and as the pressure physics. Solutions of Laplace’s equation are called harmonic
increases the velocity decreases. Bernoulli’s equation is a functions, for which there is a huge bulk of existing liter-
result of applying Newton’s second law, F = ma, to a moving ature. Therefore, it is most fortuitous that incompressible,
fluid element, as reflected in the physical derivation of the irrotational flow is described by Laplace’s equation, for which
Euler equations (Section 3). The dimensions of each term in numerous solutions exist and are well understood. In Carte-
Bernoulli’s equation are that of pressure (force per unit area). sian coordinates, Laplace’s equation is expressed as
Indeed, the second term, 21 ρV2 , is defined as the dynamic
pressure. The dimensions can also be expressed in terms of ∂2 φ ∂2 φ ∂2 φ
energy per unit volume; sometimes Bernoulli’s equation is + 2 + 2 =0 (32)
∂x2 ∂y ∂z
viewed as a type of energy equation for incompressible flow.
However, fundamentally, Bernoulli’s equation is a statement For a two-dimensional potential flow, the stream function,
of Newton’s second law. ψ, also satisfies Laplace’s equation.
For a potential flow, Bernoulli’s equation applies every-
where in the flow field and is not restricted to holding just ∂2 ψ ∂2 ψ
along a given streamline. Therefore, Bernoulli’s equation is + 2 =0 (33)
∂x2 ∂y
particularly useful for the analysis of potential flows.
From equations (31) and (33), we make the following
2.3 Divergence of the velocity obvious and important conclusions:

A mathematician would immediately think of the divergence 1. Any irrotational, incompressible flow has a velocity
of velocity in terms of its mathematical expression, which in potential and stream function (for two-dimensional flow)
Cartesian coordinates is that both satisfy Laplace’s equation
∂u ∂v ∂w 2. Conversely, any solution of Laplace’s equation rep-
∇·V = + + (29) resents the velocity potential or stream function
∂x ∂y ∂z
(two-dimensional) for an irrotational, incompressible
However, the fluid dynamicist thinks of the divergence of flow.
velocity in terms of its physical meaning. Consider a fluid
element of fixed mass moving along a streamline. In general, Laplace’s equation is a second-order linear partial dif-
the volume of this element can be changing as it moves from ferential equation. The fact that it is linear is particularly
one point to another in the flow. The divergence of velocity important, because the sum of any particular solutions of a
is physically the instantaneous time rate of change of the linear differential equation is also a solution of the equa-
volume of a moving fluid element, per unit volume. tion. For example, if φ1 , φ2 , φ3 , . . . φn represent n separation
For incompressible flow, the volume of a moving fluid solutions of equation (31), then the sum
element of fixed mass does not change. Hence, from the phys-
ical meaning of ∇·V , we have the general condition from an φ = φ1 + φ2 + . . . + φn (34)
incompressible flow that
is also a solution of equation (31). Since irrotational,
∇·V = 0 (30)
incompressible flow is governed by Laplace’s equation and
Laplace’s equation is linear, we conclude that a complicated
flow pattern for an irrotational, incompressible flow can be
2.4 Laplace’s equation synthesized by adding together a number of elementary flows
that are also irrotational and incompressible. Indeed, this
Laplace’s equation is the governing equation for incompress- establishes a grand strategy for the solution of potential flows.
ible potential flow. It stems from the condition for an incom- Develop flow-field solutions for several different elementary
pressible flow, ∇·V = 0, and the definition of the potential flows, which by themselves may not seem to be practical
function, ∇φ = V . Combining these two relations gives flows in real life. However, then proceed to add (i.e., super-
impose) these elementary flows in different ways such that
∇ 2φ = 0 (31) the resulting flow fields do pertain to practical problems.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
10 Fundamentals of Fluid Flows

Consider the irrotational, incompressible flow fields over Equation (37) or (38) gives the boundary condition for
different aerodynamic shapes, such as a sphere, cone, or velocity at the wall; it is expressed in terms of φ. If we
airplane wing. Clearly, each flow is going to be distinctly dif- are dealing with ψ rather than φ then the wall boundary
ferent; the streamlines and pressure distribution over a sphere condition is
are quite different from those over a cone. However, these dif-
ferent flows are all governed by the same equation, namely, ∂ψ
=0 (39)
∇ 2 φ = 0. How, then, do we obtain different flows for the ∂s
different bodies? The answer is found in the boundary con-
ditions. Although the governing equation for the different where s is the distance measured along the body surface.
flows is the same, the boundary conditions for the equation Note that the body contour is a streamline of the flow.
must conform to the different geometric shapes, and hence Recall that ψ = constant is the equation of a streamline.
yield different flow-field solutions. Thus, if the shape of the body is given by yb = f(x), then
Consider the external aerodynamic flow over a stationary
body. The flow is bounded by (i) the freestream flow that ψsurface = ψy=yb = constant (40)
occurs (theoretically) an infinite distance away from the body
and (ii) the surface of the body itself. Therefore, two sets of is an alternative expression for the boundary conditions
boundary conditions apply as follows: given in equation (39).

(a) Infinity boundary conditions If we are dealing with neither φ nor ψ, but rather with the
Far away from the body (toward infinity), in all velocity components u and v themselves, then the wall bound-
directions, the flow approaches the uniform freestream ary condition is obtained from the equation of a streamline
conditions. Let V∞ be aligned with the x-direction. evaluated at the body surface; that is,
Hence, at infinity
dyb v
= (41)
∂φ ∂ψ dx u surface
u= = = V∞ (35)
∂x ∂y
Equation (41) states simply that the body surface is a stream-
∂φ ∂ψ line of the flow. The form given in equation (41) for the flow
v= =− =0 (36) tangency condition at the body surface is used for all inviscid
∂y ∂x
flows, incompressible to hypersonic, and does not depend on
Equations (35) and (36) are the boundary conditions the formulation of the problem in terms of φ or ψ.
on velocity at infinity. They apply at an infinite distance
from the body in all directions, above and below, and to
the left and to right of the body. 2.5 Interim summary
(b) Wall boundary conditions
It is impossible for the flow to penetrate a solid surface. Reflecting on our previous discussions, the general approach
Instead, if the flow is viscous, the influence of friction to the solution of incompressible potential flows can be sum-
between the fluid and the solid surfaces creates a zero marized as follows:
velocity at the surface. In contrast, the velocity at the sur-
face can be finite for inviscid flows, but because the flow 1. Solve Laplace’s equation for φ (equation (32)) or ψ
cannot penetrate the surface, the velocity vector must be (equation (33)) along with the proper boundary condi-
tangent to the surface. The component of velocity nor- tions, such as equations (35) and (37). These solutions
mal to the surface must be zero. Let n be a unit vector are usually in the form of a sum of elementary solutions
normal to the surface. The wall boundary condition can (to be discussed in the following sections).
be written as 2. Obtain the flow velocity from V = ∇φ or u = ∂ψ/∂y and
v = −∂ψ/∂x.
3. Obtain the pressure from Bernoulli’s equation, p +
V ·n = (∇φ) ·n = 0 (37) 1
2
ρV 2 = p∞ + 21 ρV∞ 2 , where p∞ and V∞ are known
freestream conditions.
or

∂φ Since V and p are the primary dependent variables for an


=0 (38) incompressible flow, steps 1 to 3 are all we need to solve
∂n
Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 11

Source flow is a flow where all the streamlines are


straight lines emanating radially from a central point (the
Curve C origin) and where the flow velocity is inversely propor-
h tional to the distance r from the region. The strength of
y the source, , is the rate of volume flow from the origin,

φ = const
l
V∞ r per unit depth perpendicular to the page. The opposite
θ
x of source flow is sink flow, where the fluid flows into the
origin; a sink is essentially a source of negative strength.
ψ = const The origin of a source or sink is a mathematical singular-
ity where the flow velocity is infinite. Indeed, the origin
O can be visualized as a discrete point source or sink,
the presence of which induces the resulting radial flow
away from the origin.
Figure 2. Uniform flow.
(c) Two-dimensional doublet flow (see Figure 4)
Consider a source of strength and a sink of equal
but opposite strength – separated by a distance . Let
for a given problem as long as the flow is an incompressible the distance  approach zero while the absolute magni-
potential flow. tudes of the strength of the source and sink increase in
such a fashion that the product ( ) remains constant.
In the limit where  → 0, the resulting flow is a dou-
2.6 Elementary potential flows blet flow of strength κ =  . Again, as in the case of a
source or sink, the doublet can be considered as a discrete
Many interesting incompressible potential flows are synthe- point singularity that induces the doublet flow about the
sized from a superposition of two or more elementary flows, singularity.
all of which are solutions of Laplace’s equation. Four of the (d) Vortex flow (see Figure 5)
commonly used elementary flows are A vortex flow is a flow where all the streamlines are
concentric circles about a given point, and where the flow
(a) Uniform flow (see Figure 2) velocity is constant along any given circular streamline
This flow is simply a uniform flow moving with but varies inversely with distance from the common cen-
straight streamlines at constant velocity in a given direc- ter. This common center is called a point vortex; it is a
tion. It is the simplest of all elementary flows, and usually singularity that induces the circular vortex flow about the
constitutes the uniform freestream far ahead of an aero- common center. The circulation  (Section 1.12) taken
dynamic body. along any circular streamline about the point vortex is the
(b) Two-dimensional source flow (see Figure 3) same value, independent of which circular streamline is
t
ns
co
=
ψ

nst Vr
φ = co −Vr

r r
θ θ

O Discrete
O
Discrete sink
source

Sink flow
(a) Source flow (b)

Figure 3. Source and sink flows.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
12 Fundamentals of Fluid Flows

2.7 Equations for the elementary flows

The velocity, velocity potential, and stream function for each


of the elementary flows are derived in most aerodynamic text-
books (see, e.g., Anderson, 2010). The results are tabulated
below.
κ
2 πc
Type of flow Velocity φ ψ
Discrete doublet
Uniform flow in
κ x-Direction u = V∞ V∞ x V∞ y
Source Vr =
2␲r

2␲
ln r
2␲
θ
Vortex Vθ = − 2␲r

− 2␲

θ ln r

Doublet Vr = − 2␲
κ cos θ
r2
κ cos θ
2␲ r
− 2␲
κ sin θ
r
Vθ = − 2␲
κ sin θ
r2

2.8 Superposition of elementary flows


Figure 4. Doublet flow with strength κ.
(a) Nonlifting flow over a circular cylinder (see Figure 6)
The superposition of a uniform flow with velocity V∞

and a doublet of strength κ results in the potential flow
1
over a circular cylinder of radius R = (κ/2␲V∞ ) 2 . The
absolute magnitude of the velocity over the surface is

Vθ = −2V∞ sin θ (42)


r

and the pressure coefficient distribution over the surface


is

O p − p∞
Cp = 1 2
= 1 − 4sin2 θ (43)
ρ V
2 ∞ ∞

The pressure distribution over the surface is symmetric


about both the axes that are parallel and perpendicular to
the freestream, hence the net aerodynamic force on the
cylinder is zero.
(b) Lifting flow over a circular cylinder (see Figure 7)
The superposition of the nonlifting flow over a circu-
lar cylinder and a vortex of strength  yields a flow over
Figure 5. Vortex flows. the circular cylinder that is not symmetric about the axis
parallel to the freestream, and hence there is a net aero-
dynamic force perpendicular to the freestream, that is,
chosen. Hence,  is called the strength of the vortex. lift. The variation of pressure coefficient is given by
Vortex flow is irrotational flow everywhere except at the
origin, where the vorticity is infinity – another aspect of  2

the vortex singularity. Cp = 1 − −2sin θ − (44)
2␲RV∞

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 13

R
r

ψ=0
κ θ
+ = C D
B A

Uniform flow Doublet Flow over a cylinder


ψ = V∞ r sin θ −κ sin θ κ sin θ
ψ= ψ = V∞r sin θ −
2π r 2π r

Figure 6. Superposition of a uniform and a doublet; nonlifting flow over a circular cylinder.

By integrating the pressure coefficient over the surface, cross section when the circulation is evaluated around any
the lift per unit span perpendicular to the page is obtained curve enclosing the body. Let curve A be any curve in the flow
as enclosing the body. If the body is producing lift, the velocity
field around the body will be such that the line integral of
L = ρ∞ V∞  (45) velocity around A will be finite, that is, the circulation

Equation (45) is the Kutta–Joukowski theorem relating ⌠


lift per unit span to circulation, as discussed in the next  = − ⌡ V ·ds (46)
A
section.
is finite. In turn, the lift per unit span on the body will be given
by the Kutta–Joukowski theorem as embodied in equation
2.9 Circulation theory of lift: general philosophy (45). This result underscores the importance of the concept of
circulation. The Kutta–Joukowski theorem states that lift per
Equation (45), the Kutta–Joukowski theorem, applies in gen- unit span on a two-dimensional body is directly proportional
eral to cylindrical (i.e., two-dimensional) bodies of arbitrary to the circulation around the body.

Figure 7. The synthesis of lifting flow over a circular cylinder.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
14 Fundamentals of Fluid Flows

The lifting flow over a circular cylinder is synthesized by is obtained, then the lift per unit span follows directly from
superimposing a uniform flow, a doublet, and a vortex. All the Kutta–Joukowski theorem.
three elementary flows are irrotational at all points, except for
the vortex, which has infinite vorticity at the origin. There-
2.10 High-speed subsonic and supersonic
fore, the lifting flow over a cylinder is irrotational at every
potential flows: basic ideas and equations
point except at the origin. If we take the circulation around
any curve not enclosing the origin,  = 0. It is only when we
As long as a flow is irrotational, a velocity potential can be
choose a curve that encloses the origin, where ∇ × V is infi-
defined independent of whether the flow is incompressible or
nite, that equation (46) yields a finite , equal to the strength
compressible. Thus, for a high speed, compressible, flow, we
of the vortex. The same can be said about the potential flow
have
over any two-dimensional body. The flow outside the body is
irrotational, and the circulation around any closed curve not V = ∇φ (47)
enclosing the body is consequently zero. On the other hand,
the flow over a body is synthesized by distributing vortices where, for a two-dimensional flow,
either on the surface or inside the body. These vortices have
the usual singularities in ∇ × V, and therefore, if we choose ∂φ
u= (48)
a curve that encloses the body, equation (46) yields a finite ∂x
value of , equal to the sum of vortex strengths distributed
on or inside the body. The important point here is that, in the ∂φ
v= (49)
Kutta–Joukowski theorem, the value of  used in equation ∂y
(45) must be evaluated around a closed curve that encloses
the body. The curve can be otherwise arbitrary, but it must When the velocities in the continuity, momentum, and
have the body inside it. energy equations (Sections 1.4–1.6) are replaced by the
The approach we have discussed above – the definition of velocity potential, and suitably combined, a single equation
circulation and the use of equation (45) to obtain the lift – is results where φ is the only dependent variable.
the essence of the circulation theory of lift in aerodynamics.        
Its development at the turn of the twentieth century created a 1 ∂φ 2 ∂2 φ 1 ∂φ 2 ∂2 φ
1− 2 + 1− 2
breakthrough in aerodynamics. However, let us keep things a ∂x ∂x2 a ∂y ∂y2
in perspective. The circulation theory of lift is an alternative    2
2 ∂φ ∂φ ∂ φ
way of thinking about the generation of lift on an aerody- − 2 =0 (50)
namic body. Keep in mind that the true physical sources a ∂x ∂y ∂x∂y
of aerodynamic force on a body are the pressure and shear
stress distributions exerted on the surface of the body. The This is called the velocity potential equation. It is almost
Kutta–Joukowski theorem is simply an alternative way of completely in terms of φ; only the speed of sound appears in
expressing the consequences of the surface pressure distribu- addition to φ. However, a can be readily expressed in terms
tion; it is a mathematical expression that is consistent with the of φ as
special tools developed for the analysis of inviscid, incom-  2  2 
pressible flow. Indeed, equation (45) is derived by integrating γ −1 ∂φ ∂φ
a2 = a02 − + (51)
the pressure distribution over the surface. Therefore, it is 2 ∂x ∂y
not quite proper to say that circulation “causes” lift. Rather
lift is “caused” by the net imbalance of the surface pres- where a0 is a known constant of the flow. Equation (51) gives
sure distribution, and circulation is simply a defined quantity the speed of sound a as a function of φ. Hence, substitution
determined from the same pressures. The relation between the of equation (51) into (50) yields a single partial differential
surface pressure distribution (which produces lift L) and the equation in terms of the unknown φ. This equation repre-
circulation is given by equation (45). However, in the theory sents a combination of the continuity, momentum, and energy
of incompressible, potential flow, it is generally much easier equations. In principle, it can be solved to obtain φ for the flow
to determine the circulation around the body rather than cal- field around any two-dimensional shape, subject of course to
culate the detailed surface pressure distribution. Therein lies the usual boundary conditions at infinity and along the body
the power of the circulation theory of lift. surface.
Consequently, the theoretical analysis of lift on two- Because equation (50), along with equation (51), is a single
dimensional bodies in incompressible, inviscid flow focuses equation in terms of one dependent variable φ, the analysis of
on the calculation of the circulation about the body. Once  isentropic, irrotational, steady, compressible flow is greatly
Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 15

simplified – we only have to solve one equation instead of slightly perturbed from its original uniform flow upstream of
three or more. Once φ is known, all the other flow variables the body. Thus, the x and y components of velocity are locally
are directly obtained as follows: given by

1. Calculate u and v from equations (48) and (49). u = V∞ + û v = v̂ (55)


2. Calculate a from equation
√ (51).
3. Calculate M = V/a = u2 + v2 /a. where û and v̂ are the perturbation velocities. For a slender
4. Calculate T, p, and ρ from equations (52)–(54). body at small angle of attack, these perturbations are small
relative to the freestream velocity, that is,
T0 γ −1 2
=1+ M (52)
T 2 û v̂ û2 v̂2
, , 2, 2 ≪ 1 (56)
  γ−1
γ V∞ V∞ V ∞ V∞
p0 γ −1 2
= 1+ M (53)
p 2 With this assumption, and considering subsonic flow where
0 ≤ M∞ ≤ 0.8 or supersonic flow where 1.2 ≤ M∞ < 5
  1
ρ0 γ − 1 2 γ−1 (approximately), equation (50) reduces to
= 1+ M (54)
ρ 2
∂2 φ ∂2 φ
(1 − M∞
2
) + 2 =0 (57)
In these equations, the total conditions T0 , p0 , and ρ0 ∂x2 ∂y
are known quantities; in an isentropic flow they are con-
stant throughout the flow field and hence are obtained Equation (57) is a linear partial differential equation, and
from the freestream conditions. therefore is inherently simpler to solve than equation (50).
Since the early twentieth century, many useful analytical
Although equation (50) has the advantage of being one solutions of equation (57) have been obtained and applied
equation with one unknown, it also has the distinct disad- to subsonic and supersonic flows.
vantage of being a nonlinear partial differential equation. We have paid a price, however, for this simplicity. Equa-
Such nonlinear equations are very difficult to solve analyt- tion (57) is no longer exact. It is only an approximation to the
ically, and in modern aerodynamics, solutions of equation physics of the flow limited to cases where
(50) are usually sought by means of sophisticated numerical
techniques. Indeed, no general analytical solution of equation 1. the perturbations are small, that is, thin bodies at small
(50) has been found to this day. Contrast this situation with angles of attack;
that for incompressible flow, which is governed by Laplace’s 2. subsonic and supersonic Mach numbers. Equation (57)
equation – a linear partial differential equation for which is not valid for transonic flow, where 0.8 ≤ M∞ ≤ 1.2, or
numerous analytical solutions are well known. for hypersonic flow, where M∞ > 5.
Given this situation, aerodynamicists over the years have
made assumptions regarding the physical nature of the flow
field, which are designed to simplify equation (50). These 3 EULER EQUATIONS
assumptions limit our considerations to the flow over slender
bodies at small angles of attack. For subsonic and super- 3.1 Substantial derivative: physical meaning
sonic flows, these assumptions lead to an approximate form
of equation (50) that is linear and hence can be solved ana- The Euler equations are the governing equations for an invis-
lytically. These matters are the subject of the next section. cid flow. They are introduced in the following in both integral
Keep in mind that, within the framework of steady, irro- and partial differential equation forms. We will examine the
tational, isentropic flow, equation (50) is exact and holds for differential form in more detail further.
all Mach numbers, from subsonic to hypersonic, and for all Frequently the governing partial differential equations for
two-dimensional body shapes, thin and thick. a fluid flow are written in terms of the substantial derivative,
A linearized form of equation (50) is obtained by making which has both mathematical meaning and physical meaning.
the assumption of flow over a slender body at small angle of To introduce and understand the substantial derivative,
attack, and considering only subsonic or supersonic flow. For consider a small fluid element moving through a flow field
a slender body at small angle of attack, the flow field is only in Cartesian coordinates. The velocity field is given by

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
16 Fundamentals of Fluid Flows

V = ui + vj + wk, where derivative D/Dt. Note that Dρ/Dt is the time rate of change
of density of a given fluid element as it moves through space.
u = u(x, y, z, t) Here, our eyes are on the fluid element as it is moving, and we
v = v(x, y, z, t) (58) are watching the density of the element change as it moves
through point 1. This is different from (∂ρ/∂t)1 , which is
w = w(x, y, z, t)
physically the time rate of change of density at the fixed point
1. For (∂ρ/∂t)1 , we fix out eyes on the stationary point 1,
In addition, the density field is given by
and watch the density change due to transient fluctuations
in the flow field. Thus, Dρ/Dt and ∂ρ/∂t are physically and
ρ = ρ(x, y, z, t) (59)
numerically different quantities. Returning to equation (63),
note that
At time t1 , the fluid element is located at point 1 in the flow,
and its density is x2 − x 1
lim ≡u
t2 →t1 t2 − t 1
ρ1 = ρ(x1 , y1 , z1 , t1 ) (60) y2 − y 1
lim ≡ v (65)
t2 →t1 t2 − t1
At a later time, t2 the same fluid element has moved to a z2 − z1
different location in the flow field, such as point 2. At this lim ≡w
t2 →t1 t2 − t1
new time and location, the density of the fluid element is
Thus, taking the limit of equation (63) as t2 → t1 , we obtain
ρ2 = ρ(x2 , y2 , z2 , t2 ) (61)
Dρ ∂ρ ∂ρ ∂ρ ∂ρ
=u +v +w + (66)
Since ρ = ρ(x, y, z, t), we can expand this function in a Taylor Dt ∂x ∂y ∂z ∂t
series about point 1 as follows:
    From equation (66), we can obtain an expression for the
∂ρ ∂ρ substantial derivative in Cartesian coordinates:
ρ2 = ρ1 + (x2 − x1 ) + (y2 − y1 )
∂x 1 ∂y 1
  D ∂ ∂ ∂ ∂
+
∂ρ
(z2 − z1 ) (62) = +u +v +w (67)
∂z 1 Dt ∂t ∂x ∂y ∂z
 
∂ρ
+ (t2 − t1 ) + higher order terms Furthermore, in Cartesian coordinates, the vector operator ∇
∂t 1 is defined as
Dividing by t2 − t1 , and ignoring the higher order terms, we ∂ ∂ ∂
have ∇≡i +j +k (68)
∂x ∂y ∂z
   
ρ 2 − ρ1 ∂ρ x2 − x1 ∂ρ x2 − x1
= + Hence, equation (67) can be written as
t2 − t 1 ∂x 1 t2 − t1 ∂y 1 t2 − t1
   
∂ρ z2 − z1 ∂ρ D ∂
+ + (63) ≡ + (V ·∇) (69)
∂z 1 t2 − t1 ∂t 1 Dt ∂t

Consider the physical meaning of the left side of equation Equation (69) represents a definition of the substantial deriva-
(63). The term (ρ2 − ρ1 )/(t2 − t1 ) is the average time rate of tive in vector notation; thus, it is valid for any coordinate
change in density of the fluid element as it moves from point 1 system.
to point 2. In the limit, as t2 approaches t1 , this term becomes Focusing on equation (69), we once again emphasize that
D/Dt is the substantial derivative, which is physically the
ρ2 − ρ 1 Dρ time rate of change following a moving fluid element; ∂/∂t
lim = (64)
t2 →t1 t2 − t 1 Dt is called the local derivative, which is physically the time
rate of change at a fixed point; V ·∇ is called the convective
Here, Dρ/Dt is a symbol for the instantaneous time rate of derivative, which is physically the time rate of change due
change of density of the fluid element as it moves through to the movement of the fluid element from one location to
point 1. By definition, this symbol is called the substantial another in the flow field where the flow properties are spatially

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 17

different. The substantial derivative applies to any flow-field Momentum:


variable (e.g., Dp/Dt, Dt/Dt, Du/Dt). For example,
Du ∂p
ρ = − + ρfx (72)
Dt ∂x
DT ∂T ∂T ∂T ∂T ∂T
≡ + (V ·∇)T ≡ +u +v +w
Dt ∂t ∂t ∂x ∂y ∂z Dv ∂p
Local
Convective ρ = − + ρfy (73)
derivative
derivative Dt ∂y

(70) Dw ∂p
ρ = − + ρfz (74)
Dt ∂z
Equation (70) states physically that the temperature of the
fluid element is changing as the element sweeps past a point Energy:
in the flow because at that point the flow-field temperature
itself may be fluctuating with time (the local derivative) and D(e + V2
)
because the fluid element is simply on its way to another ρ 2
= ρq̇ − ∇·(pV ) + ρ(f ·V ) (75)
Dt
point in the flow field where the temperature is different (the
convective derivative). The above continuity, momentum, and energy equations are
Consider an example that will help to reinforce the phys- three equations in terms of the five unknowns p, V, ρ, T,
ical meaning of the substantial derivative. Imagine that you and e. Assuming a calorically perfect gas, the additional two
are hiking in the mountains, and you are about to enter a cave. equations needed to complete the system are obtained from
The temperature inside the cave is cooler than outside. Thus,
as you walk through the mouth of the cave, you feel a temper- Equation of state: p = ρRT
ature decrease – this is analogous to the convective derivative Internal energy: e = cv T
in equation (70). However, imagine that, at the same time, a
friend throws a snowball at you such that the snowball hits
you just at the instant you pass through the mouth of the
3.3 Boundary conditions
cave. You will feel an additional, but momentary, temper-
ature drop when the snowball hits you – this is analogous
The Euler equations describe an inviscid flow. The boundary
to the local derivative in equation (70). The net temperature
conditions dictate the particular flow being studied. Many
drop you feel as you walk through the mouth of the cave is
problems in aerodynamics deal with the external flow over a
therefore a combination of both the act of moving into the
body of given shape. For these problems, there are two sets
cave, where it is cooler, and being struck by the snowball at
of boundary conditions.
the same instant – this net temperature drop is analogous to
Uniform freestream: At the appropriate distance from the
the substantial derivative in equation (70).
body, the flow is treated as uniform with given freestream
conditions
3.2 Euler equations in terms of the substantial
derivative V ∞ = constant p∞ = constant ρ∞ = constant (76)

On a classical basis, the term Euler equations applies to Body surface: The flow cannot penetrate the surface.
the momentum equation for an inviscid flow. However, the Therefore, at the body, the flow must be locally tangent to
modern use of the term Euler equations covers the whole the surface. If n is a unit normal vector perpendicular to the
system of governing fundamental equations for an inviscid surface, then
flow, the continuity, momentum, and energy equations. These
equations are introduced in Brief History of the Early Devel-
V ∞ ·n = 0 (77)
opment of Theoretical and Experimental Fluid Dynamics.
Written in terms of the substantial derivative, these equations
are
Continuity: 3.4 Discussion of the Euler equations

Dρ The Euler equations are a coupled system of nonlinear partial


+ ρ∇·V = 0 (71) differential equations, in the form of the substantial derivative
Dt

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
18 Fundamentals of Fluid Flows

given in this chapter. To date, no general closed-form analyt- x-,y-, and z-directions, expressed in terms of the substantial
ical solution to these equations has been found. Therefore, derivative (Section 3) is
two approaches to their solution are

1. Simplify the physical problem by using appropriate Du ∂u ∂u ∂u ∂u


approximations that lead to a linearized system of equa- ρ = ρ + ρu + ρv + ρw
Dt ∂t ∂x ∂y ∂z
tions that can be solved numerically. Various small  
∂p ∂ ∂u
perturbation solutions suitable for a slender body at small =− + λ∇·V + 2µ (78)
∂x ∂x ∂x
angle of attack, and for subsonic or supersonic flows, are      
∂ ∂v ∂u ∂ ∂u ∂w
an example. + µ + + µ + + ρfx
2. Carry out a numerical solution of the full Euler equations ∂y ∂x ∂y ∂z ∂z ∂x
using an appropriate algorithm from computational fluid
dynamics (CFD). Today, such numerical solutions of the
Euler equations are the norm in aerodynamics. Dv ∂v ∂v ∂v ∂v
ρ = ρ + ρu + ρv + ρw
Dt ∂t ∂x ∂y ∂z
  
∂p ∂ ∂v ∂u
=− + µ + (79)
3.5 Physical significance of solutions of the Euler ∂y ∂x ∂x ∂y
    
equations ∂ ∂v ∂ ∂w ∂v
+ λ∇·V + 2µ + µ + + ρfy
∂y ∂y ∂z ∂y ∂z
Today, CFD solutions of the Euler equations are carried out
for the flow over virtually any complex shape, such as a com-
plete three-dimensional airplane at any orientation to the flow.
Dw ∂w ∂w ∂w ∂w
These solutions apply to any flow that is not influenced by the ρ =ρ + ρu + ρv + ρw
Dt ∂t ∂x ∂y ∂z
effects of friction. Fully attached flows over bodies are readily   
∂p ∂ ∂u ∂w
handled by the Euler equations. Solutions of the Euler equa- =− + µ + (80)
tions yield reasonably valid results for the surface pressure ∂z ∂x ∂z ∂x
    
distribution; since lift and moments on a body are due to the ∂ ∂w ∂v ∂ ∂w
+ µ + + λ∇·V + 2µ + ρfz
pressure distribution integrated over the body surface, these ∂y ∂y ∂z ∂z ∂z
aerodynamic forces and moment are usually accurately pre-
dicted by CFD Euler solutions. On the other hand, solutions
of the Euler equations provide no direct results for aerody- where µ is the coefficient of viscosity and λ is the second
namic drag, which is due to shear stress at the surface, and to coefficient of viscosity.
regions of flow separation over the body. Solutions for drag Equations (78)–(80) represent the complete Navier–
are the purview of either Navier–Stokes solutions (Section 4) Stokes equations for an unsteady, compressible, three-
or boundary-layer solutions (Section 5). dimensional viscous flow. To analyze incompressible viscous
flow, equations (78)–(80) and the continuity equation are
sufficient. However, for a compressible flow, we need an addi-
4 NAVIER–STOKES EQUATIONS tional equation, namely, the energy equation to be discussed
next.
4.1 Expanded form of the Navier–Stokes
equations

4.1.1 Momentum equation 4.1.2 Energy equation


Applying the flow model of an infinitesimally small mov- In Section 1.6, the viscous terms in the energy equation,
ing fluid element of fixed mass (see Section 2), the viscous Q̇1viscous and Ẇviscous
1
, were treated generically. These terms
forces in the momentum equation are obtained by treating are due to thermal conduction and viscous stresses acting
the viscous shear and normal stresses acting on all sides of on a fluid element, and are functions of the temperature and
the fluid element. In turn, these viscous stresses are related to velocity gradients, respectively, in the flow. In these terms, the
appropriate velocity gradients in the flow (see, e.g., Ander- energy equation written in terms of the substantial derivative
son, 2010). The final form of the momentum equations in the is

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 19

      the following dimensionless variables:


De ∂ ∂T ∂ ∂T ∂ ∂T
ρ = ρq̇ + k + k + k
Dt ∂x ∂x ∂y ∂y ∂z ∂z ρ u v p
    ρ = u = v = p =
∂u ∂v ∂w ∂u ∂v ∂w 2
−p + + +λ + + ρ∞ V∞ V∞ p∞
∂x ∂y ∂z ∂x ∂y ∂z
    2  2 µ x y
∂u 2
∂v ∂w µ = x = y = (83)
+µ 2 +2 +2 µ∞ c c
∂x ∂y ∂z
 2  2    where ρ∞ , v∞ , p∞ , and µ∞ are reference values (say, e.g.,
∂u ∂v ∂u ∂w ∂v ∂w 2
+ + + + + + freestream values) and c is a reference length (say, the chord
∂y ∂x ∂z ∂x ∂z ∂y of an airfoil). In terms of these dimensionless variables, equa-
(81) tion (82) becomes
  
∂u ∂u p∞ ∂p
where e is the internal energy and k is the thermal conductiv- ρ  u 
+ ρ  v  = −
ity. Please note that the energy equation can be written in a ∂x ∂y 2
ρ∞ V ∞ ∂x
     
number of different forms; equation (81) is just one of them µ∞ ∂  ∂v ∂u
dealing with the internal energy. + µ +
ρ∞ V∞ c ∂y ∂x ∂y
(84)
4.2 Similarity parameters
Noting that
One of the most important concepts in fluid dynamics is that
of flow-field dynamic similarity. Two different flows exhibit p∞ γp∞ 2
a∞ 1
flow-field dynamic similarity if the variation of nondimen- 2
= 2
= 2
= 2
(85)
ρ∞ V ∞ γρ∞ V∞ γV∞ γM∞
sional dependent variables as a function of nondimensional
independent variables is the same for both flows. To ensure and
flow-field similarity, certain parameters called similarity
parameters must be the same between the two flows. The µ∞ 1
= (86)
most rigorous method of identifying these similarity param- ρ∞ V∞ c Re∞
eters is to nondimensionalize the Navier–Stokes and energy
equations, and examine the coefficients of the nondimen- where M∞ and Re∞ are the freestream Mach and Reynolds
sional terms (see, e.g., Anderson 2010). These similarity numbers, respectively, equation (84) becomes
parameters are the Mach numbers, M∞ = Va∞∞ , Reynolds
µ c
number, Re = ρ∞µV∞∞ c , Prandtl number, Pr = k∞∞ p , and the ∂u ∂u 1 ∂p
cp
ρ  u 
+ ρ  v  = − 2 ∂x
ratio of specific heats γ = cv . If the similarity parameters are ∂x ∂y γM∞
   
the same for two different flows, the two flows are dynami- 1 ∂  ∂v ∂u
cally similar. In these parameters, c is a characteristic length, + µ +
Re∞ ∂y ∂x ∂y
cp and cv are the specific heats at constant pressure and con-
stant volume respectively, a∞ is the speed of sound, ρ∞ is (87)
the density, V∞ is the freestream velocity, µ∞ is the viscosity
coefficient, and k∞ is the thermal conductivity. Equation (87) tells us something important. Consider two dif-
To illustrate this process of nondimensionalizing the ferent flows over two bodies of different shapes. In one flow,
governing equations, consider a steady two-dimensional, vis- the ratio of specific heats, Mach number, and Reynolds num-
cous, compressible flow. The x-momentum equation for such ber are γ1 , M∞1 , and Re∞1 , respectively; in the other flow,
a flow is given by equation (78), which for the present case these parameters have different values, γ2 , M∞2 , and Re∞2 .
reduces to Equation (87) is valid for both flows. It can, in principle, be
solved to obtain u as a function of x and y . However, since
  
∂u ∂u ∂p ∂ ∂v ∂u γ, M∞ , and Re∞ are different for the two cases, the coeffi-
ρu + ρv =− + µ + (82) cients of the derivatives in equation (87) will be different.
∂x ∂y ∂x ∂y ∂x ∂y
This will ensure, if
In equation (82), ρ, u, p, and so on are the actual dimen-
sional variables, say [ρ] = kg/m3 , and so on. Let us introduce u = f1 (x , y ) (88)

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
20 Fundamentals of Fluid Flows

represents the solution for one flow and For the energy equation, the surface boundary condition
can be treated as any of the following. If the wall temperature
u = f2 (x , y ) (89) Tw is known, then the boundary condition at the surface is
simply
represents the solution for the other flow, that
T = Tw (at the wall) (94)
f1 =
/ f2 (90)
On the other hand, if the wall temperature is not known,
However, consider now the case where the two different for example, if it is changing as a function of time due to aero-
flows have the same values of γ, M∞ , and Re∞ . Now the coef- dynamic heat transfer to or from the surface, then the Fourier
ficients of the derivatives in equation (87) will be the same for law of heat conduction provides the boundary condition at
both flows; that is, equation (87) is numerically identical for the surface. If q̇w denotes the instantaneous heat flux to the
the two flows. In addition, assume the two bodies are geomet- wall, then
rically similar, so that the boundary conditions in terms of the
 
nondimensional variables are the same. Then, the solutions ∂T
of equation (87) or the two flows in terms of u = f1 (x , y ) q̇w = −k (at the wall) (95)
∂n w
and f2 (x , y ) must be identical; that is,
where n denotes the direction normal to the wall. Here, the
f1 (x , y ) ≡ f2 (x , y ) (91) surface material is responding to the heat transfer to the wall,
q̇w , hence changing Tw , which in turn affects q̇w . This general
Recall the definition of dynamically similar flows given unsteady heat-transfer problem must be solved by treating the
earlier, namely, that two flows are dynamically similar if viscous flow and the thermal response of the wall material
the distributions of V/V∞ , p/p∞ , and so on are the same simultaneously. This type of boundary condition, as far as
throughout the flow field when plotted against common the flow is concerned, is a boundary condition on the temper-
nondimensional coordinates. This is precisely what equation ature gradient at the wall, in contrast to stipulating the wall
(91) is addressing – that u as a function of x and y is the same temperature itself as the boundary condition. That is,
for the two flows. That is, the variation of the nondimensional
 
velocity as a function of the nondimensional coordinates ∂T q̇w
is the same for the two flows. How did we obtain equation =− (96)
∂n w k
(91)? Simply by saying that γ, M∞ , and Re∞ are the same
for the two flows and that the two bodies are geometrically Finally, when the wall temperature becomes such that
similar. These are precisely the criteria for two flows to be there is no heat transfer to the surface, this wall tempera-
dynamically similar. ture, by definition, is called the adiabatic wall temperature,
Taw . Hence, for an adiabatic wall, the boundary condition is

4.3 Boundary conditions  


∂T
=0 (at the wall) (97)
∂n w
The boundary conditions for the Navier–Stokes and energy
equations are completely different from those for the Euler
equations (Section 3), except for the uniform freestream.
Uniform freestream: At the appropriate distance from the
body, the flow is treated as uniform with given freestream 4.4 Discussion of the Navier–Stokes equations
conditions.
The Navier–Stokes and energy equations are a coupled sys-
V∞ = constant p∞ = constant ρ∞ = constant (92) tem of nonlinear partial differential equations, even more
complex than the Euler Equations discussed in Section 3.
Body surface: In contrast to the case of the Euler equations To date, no general closed-form analytical solution to these
(Section 3), in a viscous flow the no-slip condition holds at equations has been found. Therefore, two approaches to their
the surface, that is, the flow velocity relative to the body at solution are
the surface is zero:
1. simplify the physical problem by using appropriate
u=0 v=0 w=0 (93) approximations that lead to simpler mathematics or
Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 21

2. carry out a numerical solution for the full equations using Navier–Stokes equations (Section 4), and of course such
an appropriate algorithm from CFD. Such numerical solutions were unobtainable. In 1904, one of the major
solutions of the Navier–Stokes equations are the norm breakthroughs in fluid dynamics occurred when Ludwig
in aerodynamics (see Anderson, 1995). Prandtl advanced his concept of the boundary layer. He rea-
soned that the region of the flow where viscous effect would
be important is a thin layer immediately adjacent to a sur-
4.5 Physical significance of solutions of the
face – the boundary layer – and that the remainder of the
Navier–Stokes equations
flow external to the boundary layer could be treated as an
inviscid flow (see Brief History of the Early Development
Today, with some intensive effort and great computer power,
of Theoretical and Experimental Fluid Dynamics and Sec-
CFD solutions are carried out for flow over many complex
tion 2). The equations that govern the viscous flow in the
body shapes. If the flow involves regions of flow separa-
boundary layer are still nonlinear partial differential equa-
tion, then a numerical solution of the complete Navier–Stokes
tions, but they are of a mathematical nature that makes them
equations is virtually the only choice. These solutions pro-
easier to solve than the Navier–Stokes equations, that is,
vide both the pressure and the shear stress distributions over
the boundary layer equations are parabolic partial differen-
a body surface; hence, they can be used to predict both lift
tial equations that lend themselves directly to a downstream
and drag on the body. However, the drag predictions are
marching solution whereas the Navier–Stokes equations have
very sensitive to how the CFD solution handles turbulent
a mixed elliptic–parabolic nature that precludes a straightfor-
flow. To date, such solutions have to model the influence
ward downstream marching solution.
of turbulence, and these different turbulence models are fre-
quently the Achilles heel for drag predictions. In the extreme,
if a detailed enough numerical grid with very small spac-
5.2 Boundary-layer properties
ing is used, the Navier–Stokes solutions can actually predict
and simulate the turbulence itself. These solutions are called
For simplicity, consider the viscous flow over a flat plate as
direct numerical simulations (DNS), but such DNS solutions
shown in Figure 8.
require vast amounts of computer power and memory so
The viscous effects are contained within a thin layer adja-
as to make them (as of today) impractical for engineering
cent to the surface; the thickness is exaggerated as shown
calculations of real problems.
in Figure 8 for clarity. Immediately at the surface, the
flow velocity is zero; this is the “no-slip” condition dis-
5 BOUNDARY-LAYER EQUATIONS cussed in Section 4. In addition, the temperature of the fluid
immediately at the surface is equal to the temperature of the
5.1 Background surface; this is called the wall temperatureTw , as shown in
Figure 8. Above the surface, the flow velocity increases in
Throughout the nineteenth century, physicists and mathe- the y-direction until, for all practical purposes, it equals the
maticians felt that the only way to properly account for freestream velocity. This will occur at a height above the
viscous effects in a fluid flow was to solve the complete wall equal to δ, as shown in Figure 8. More precisely, δ is

,T = Te
ry layer
unda
V∞
er mal bo
h
e of t
T∞ r edg
Oute u = ue
ry layer,
f velocit y bounda
r edge o δT
Oute T
y
δ
Vw = 0,T = Tw Tw
x


qw Velocity Temperature
profile profile

Figure 8. Boundary-layer properties.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
22 Fundamentals of Fluid Flows

defined as the distance above the wall where u = 0.99ue ; c


ody δ∗
here, ue is the velocity at the outer edge of the boundary layer. tive b
Effec b
In Figure 8, which illustrates the flow over a flat plate, the ce
dy surfa
velocity at the edge of the boundary layer will be V∞ ; that Actual bo
V∞
is, ue = V∞ . For a body of general shape, ue is the veloc- a
ity obtained from an inviscid flow solution evaluated at the
body surface (or at the “effective body” surface, as discussed
later). The quantity δ is called the velocity boundary-layer
thickness. At any given x station, the variation of u between
y = 0 and y = δ, that is, u = u(y), is defined as the veloc-
Figure 9. The “effective body,” equal to the actual body shape plus
ity profile within the boundary layer, as shown in Figure 8. the displacement thickness distribution.
This profile is different for different x stations. Similarly,
the flow temperature will change above the wall, ranging
from T = Tw at y = 0 to T = 0.99Te at y = δT . Here, δT q̇w = q̇w (x). One of the central purposes of boundary-layer
is defined as the thermal boundary-layer thickness. At any theory is to compute τw and q̇w .
given x station, the variation of T between y = 0 and y = δT , A frequently used boundary-layer property is the displace-
that is, T = T (y), is called the temperature profile within the ment thickness δ∗ , defined as
boundary layer, as shown in Figure 8. (In the above, Te is the
temperature at the edge of the thermal boundary layer. For y1  
ρu
the flow over a flat plate, as shown in Figure 8, Te = T∞ . δ∗ ≡ 1− dy δ ≤ y1 → ∞ (100)
0 ρe ue
For a general body, Te is obtained from an inviscid flow
solution evaluated at the body surface, or at the “effective
body” surface, to be discussed later.) Hence, two boundary The practical use of δ∗ is to define an “effective body shape”
layers can be defined: a velocity boundary layer with thick- given by the actual body contour plus the displacement thick-
ness δ and a temperature boundary layer with thickness δT . ness, as shown in Figure 9.
In general, δT = / δ. The relative thicknesses depend on the Here, the body is given by curve ab. However, due to the
Prandtl number; it can be shown that if Pr = 1, then δ = δT ; displacement effect of the boundary layer, the shape of the
if Pr > 1, then δT < δ; if Pr < 1, then δT > δ. For air at body effectively seen by the freestream is not given by curve
standard conditions, Pr = 0.71; hence, the thermal boundary ab; rather, the freestream sees an effective body given by
layer is thicker than the velocity boundary layer, as shown in curve ac. In order to obtain the conditions ρe , ue , Te , and so
Figure 8. Note that both boundary-layer thicknesses increase on at the outer edge of the boundary layer on the actual body
with distance from the leading edge; that is δ = δ(x) and ab, an inviscid flow solution should be carried out for the
δT = δT (x). effective body, and ρe , ue , Te , and so on are obtained from
The consequence of the velocity gradient at the wall is the this inviscid solution evaluated along curve ac.
generation of shear stress at the wall, In many cases, the boundary layers are so thin that the
effective body can be ignored, and a boundary-layer solu-
 
∂u tion can proceed directly from ρe , ue , Te , and so on obtained
τw = µ (98) from an inviscid solution evaluated on the actual body sur-
∂y w
face (ab in Figure 9). However, for highly accurate solutions,
 
and for cases where the boundary-layer thickness is rel-
where ∂u∂y w
is the velocity gradient evaluated at y = 0 (i.e.,
atively large (such as for hypersonic flow), the iterative
at the wall). Similarly, the temperature gradient at the wall procedure described above should be carried out. Also, we
generates heat transfer at the wall, note parenthetically that δ∗ is usually smaller than δ; typically
  δ∗ ≈ 0.3δ.
∂T
q̇w = −k (99)
∂y w

  5.3 The boundary-layer equations


where ∂T ∂y w
is the temperature gradient evaluated at y = 0
(i.e., at the wall). In general, both τw and q̇w are functions For simplicity, we consider two-dimensional steady flow. The
of distance from the leading edge; that is, τw = τw (x) and nondimensionalized form of the x-momentum equation (one

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 23

(104), we have
δ≪c
V∞
y [O(1)][O(1)] [O(1)][v ]
δ + =0 (105)
x O(1) O(δ)

c Hence, from equation (105), clearly v must be of an order


of magnitude equal to δ; that is, v = O(δ). Now examine the
Figure 10. The basic assumption of boundary-layer theory: A order of magnitude of the terms in equation (101). We have
boundary layer is very thin in comparison with the scale of the
body. ∂u ∂u ∂p
ρ  u = O(1) ρ  v = O(1) = O(1)
∂x ∂y ∂x
of the Navier–Stokes equations) was developed in Section 4      
∂  ∂v ∂  ∂u 1
as µ = O(1) µ = O (106)
∂y ∂x ∂y  ∂y  δ2
∂u   ∂u

1 ∂p
ρ  u + ρ v = − Hence, the order-of-magnitude equation for equation (101)
∂x ∂y 2 ∂x
γM∞
    can be written as
1 ∂  ∂v ∂u   
+ µ + 1 1 1
Re∞ ∂y ∂x ∂y O(1) + O(1) = − O(1) + O(1) + O 2
γM∞ 2 Re∞ δ
(101)
(107)
We now reduce equation (101) to an approximate form that
holds reasonably well within a boundary layer. Let us now introduce another assumption of boundary-layer
Consider the boundary layer along a flat plate of length c as theory, namely, the Reynolds number is large, indeed, large
shown in Figure 10. The basic assumption of boundary-layer enough such that
theory is that a boundary layer is very thin in comparison
1
with the scale of the body; that is, = O(δ2 ) (108)
Re∞
δ
c (102)
Then equation (106) becomes

Consider the continuity equation for a steady, two-   


1 1
dimensional flow, O(1) + O(1) = − 2
O(1) + O(δ ) O(1) + O 2
2
γM∞ δ
∂ (ρu) ∂ (ρv) (109)
+ =0 (103)
∂x ∂y
In equation (109), there is one term with an order of
In terms of the nondimensional variables defined in Section magnitude that is much smaller than the rest, namely,
4, equation (103) becomes the product
   O(δ  
2
)[O(1)] = O(δ2 ). This term corresponds to
1 ∂
µ ∂x
∂v
in equation (101). Hence, neglect this term


Re∞ ∂y 

∂ ρ  u ∂ ρ  v in comparison to the remaining terms in equation (101). We


+ =0 (104)
∂x ∂y obtain
   
Because u varies from 0 at the wall to 1 at the edge of the   ∂u   ∂u 1 ∂p 1 ∂  ∂u
ρu  +ρv  =− 2 ∂x
+ µ 
boundary layer, let us say that u is of the order of magni- ∂x ∂y γM∞ Re∞ ∂y ∂y
tude equal to 1, symbolized by O(1). Similarly, ρ = O(1). (110)
Also, since x varies from 0 to c, x = O(1). However, since y
varies from 0 to δ, where δ
c, then y
is of the smaller In terms of dimensional variables, equation (110) is given by
order of magnitude, denoted by y = O cδ . Without loss
 
of generality, we can assume that c is a unit length. There- ∂u ∂u ∂p ∂ ∂u
ρu + ρv =− + µ (111)
fore, y = O(δ). Putting these orders of magnitude in equation ∂x ∂y ∂x ∂y ∂y

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
24 Fundamentals of Fluid Flows

Equation (111) is the approximate x-momentum equation that ∂p


y-Momentum : =0 (118)
holds for flow in a thin boundary layer at a high Reynolds ∂y
number.
Consider the y-momentum equation for two-dimensional,
steady flow (Section 4):  
∂h ∂h ∂ ∂T dpe
   Energy : ρu + ρv = k +u
∂v ∂v ∂p ∂ ∂v ∂u ∂x ∂y ∂y ∂y dx
ρu + ρv = − + µ + (112)  2
∂x ∂y ∂y ∂x ∂x ∂y ∂u
+µ (119)
∂y
In terms of the nondimensional variables, equation (112)
becomes
Note that, as in the case of the Navier–Stokes equations,
∂v ∂v 1 ∂p the boundary-layer equations are nonlinear. However, the
ρ  u  + ρ  v  = − 2 ∂y boundary-layer equations are simpler, and therefore are more
∂x ∂y γM∞
    readily solved. Also, since p = pe (x), the pressure gradient
1 ∂  ∂v ∂u expressed as ac∂p∂x is reexpressed as dp e
in equations (117)
+ µ + dx
Re∞ ∂x ∂x ∂y and (119). In the above equations, the unknowns are u, v, ρ,
(113) and h; p is known from p = pe (x), and µ and k are prop-
erties of the fluid which vary with temperature. To complete
The order-of-magnitude equation for equation (113) is the system, we have

  
1 ∂p 1 p = ρRT (120)
O(δ) + O(δ) = − 2 ∂y
+ O(δ ) O(δ) + O
2
γM∞ δ
(114) h = cp T (121)


From equation (114), we see that ∂p
∂y
= O(δ) or smaller, Hence, equations (116), (117), and (119) to (121) are five
assuming that γM∞ = O(1). Since δ is very small, this
2 equations for the five unknowns, u, v, ρ, T , and h.

implies that ∂p
∂y
is very small. Therefore, from the y- The boundary conditions for the above equations are as
momentum equation specialized to a boundary layer, we follows:
have
At the wall : y=0 u=0 v=0 T = Tw
∂p
=0 (115) At the boundary-layer edge: (122)
∂y
y→∞ u → ue T → Te
Equation (113) is important; it states that at a given x-
station, the pressure is constant through the boundary layer Note that since the boundary-layer thickness is not known
in a direction normal to the surface. This implies that the a priori, the boundary condition at the edge of the boundary
pressure distribution at the outer edge of the boundary layer layer is given at large y, essentially y approaching infinity.
is impressed directly to the surface without change. Hence,
throughout the boundary layer, p = p(x) = pe (x).
The energy equation (Section 4) can be treated in the same 5.4 Solutions of the boundary-layer equations
fashion. In summary, the boundary-layer equations are
The boundary-layer equations are nonlinear partial differen-
∂(ρu) ∂(ρv)
Continuity : + =0 (116) tial equations that have to be solved numerically. However,
∂x ∂y because they are parabolic partial differential equations, they
generally lend themselves to a straightforward downstream
marching solution. The solution of these equations yields the
x-Momentum : velocity and temperature profiles throughout the boundary
 
∂u ∂u dpe ∂ ∂u layer, including at the wall. However, the practical informa-
ρu + ρv =− + µ (117)
∂x ∂y dx ∂y ∂y tion we want is the solution for τw and q̇w , the surface shear

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 25

stress and heat transfer, respectively. These are given by 1. The thermodynamic properties (e, h, p, T , ρ, s, etc.) are
completely different.
 
τw = µw ∂u 2. The transport properties (µ and k) are completely dif-
∂y w
  (123) ferent. Moreover, the additional transport mechanism of
and q̇w = −kw ∂T
∂y w
diffusion becomes important, with the associated diffu-
sion coefficients Di,j .
Therefore, in order to obtain the surface shear stress and 3. High heat-transfer rates are usually a dominant aspect of
heat transfer, we should have the velocity gradient and the any high-temperature application.
temperature gradient at the wall. These gradients in turn can 4. The ratio of specific heats, γ = cp /cv , is a variable.
only be obtained by a complete solution of the velocity and In fact, for the analysis of high-temperature flows, γ
temperature profiles throughout the complete boundary. loses the importance it has for the classical constant γ
flows. From this point of view, all equations derived
under the assumption of a constant γ are not valid for a
high-temperature gas. Such equations represent the vast
6 CHEMICALLY REACTING FLOW majority of classical aerodynamic analysis. In the pro-
EQUATIONS cess, we lose the ability for closed-form analyses using
such equations.
6.1 Importance of chemically reacting flows: 5. In view of the preceding, virtually all analyses of high-
hypersonic flight and combustion temperature gas flows require some type of numericals,
rather than closed-form, solutions.
On 24 July 1969, Apollo 11 successfully entered the atmo- 6. If the temperature is high enough to cause ionization,
sphere of the Earth, returning from its historic first manned the gas becomes a partially ionized plasma, which has a
flight to the moon. During its return to Earth, the Apollo finite electrical conductivity. In turn, if the flow is in the
vehicle acquired a velocity essentially equal to escape veloc- presence of an exterior electric or magnetic field, then
ity from the Earth, approximately 11.2 km s−1 . At this entry electromagnetic body forces act on the fluid elements.
velocity, the shock layer temperature (the flow between the This is the purview of an area called magnetohydrody-
curved detached bow shock wave and the body) becomes namics (MHD).
very large. Using the methods from classical compressible 7. If the gas temperature is high enough, there will be nona-
flow, we can predict the gas temperature in the shock layer diabatic effects caused by radiation to or from the gas.
to reach 58 128 K – ungodly high, but also totally incorrect.
The classical calculation assumes a nonreacting gas with a
constant value of the ratio of specific heats γ = 1.4. In reality, 6.3 Chemical effects in high-temperature air
at such high temperatures air becomes chemically reacting,
and γ no longer equals 1.4 nor is it constant. A more realistic At what temperatures do chemically reacting effects become
calculation, assuming the flow to be in local chemical equi- important in air? Imagine that we take the air in the room
librium shows the shock layer temperature to be 11 600 K – around us and progressively increase the temperature, hold-
also a very high temperature but considerably lower than ing the pressure constant at 1 atm. At about a temperature
the 58 128 K originally predicted. The major points here are of 800 K, the vibrational energy of the molecules becomes
(i) the temperature in the shock layer of a high-speed entry significant (as shown in Figure 11, right). This is not a chem-
vehicle can be very high and (ii) if this temperature is not cal- ical reaction, but it does have some impact on the properties
culated properly, huge errors will result. Therefore, the proper of the gas. When the temperature reaches about 2000 K, the
treatment of high temperature chemically reacting flows is dissociation of O2 begins. At 4000 K, the O2 dissociation
essential for many high-speed aerodynamic applications, and is essentially complete; most of the oxygen is in the form of
in combustion phenomena. atomic oxygen O. Moreover, by an interesting quirk of nature,
4000 K is the temperature at which N2 begins to dissociate, as
shown in Figure 11. When the temperature reaches 9000 K,
6.2 Nature of chemically reacting flows most of the N2 has dissociated. Coincidentally, this is the tem-
perature at which both oxygen and nitrogen ionization occurs,
What in general makes high-temperature flows any different and above 9000 K, we have a partially ionized plasma con-
to study and calculate than the flow of a gas with constant γ? sisting mainly of O, O+ , N, N+ , and electrons. Not shown in
The answer is as follows: Figure 11 (because it would become too cluttered) is a region

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
26 Fundamentals of Fluid Flows
Range of ionization

Repulsive
N N+ + e −
force
O O+ + e −
(+)
9000 K N2 almost completely dissociated;
ionization begins
Range of dissociation

N2 2N Distance from molecule

N2 begins to dissociate; O2 is
4000 K almost completely dissociated
O2 2O Attractive Force field is
2500 K O2 begins to dissociate force usually negligible,
Vibrational (−) about 10 molecular
No reactions 800 K
excitation diameters away from
0K the molecule

Figure 11. Ranges of vibrational excitation, dissociation, and ion-


ization for air at 1-atm pressure. Figure 12. Sketch of the intermolecular force variation.

of mild ionization that occurs around 4000–6000 K; here, one of these molecules out of the air around you. Exam-
small amounts of NO are formed, some of which ionize to ine it closely. You will find that a force field surrounds this
form NO+ and free electrons. In terms of the overall chemical molecule, as a result of the electromagnetic action of the
composition of the gas, these are small concentrations; how- electrons and nuclei of the molecule. In general, this force
ever, the electron number density as a result of NO ionization field will reach out from the given molecule and will be felt
can be sufficient to cause a communications blackout to and by neighboring molecules, and vice versa. Thus, the force
from an entry vehicle. From Figure 11, it is very useful to fix in field is called an intermolecular force. A schematic of a typ-
mind the “onset” temperatures: 800 K for vibrational excita- ical intermolecular force field caused by a single particle is
tion, 2500 K for O2 dissociation, 4000 K for N2 dissociation, shown in Figure 12. Here, the intermolecular force is sketched
and 9000 K for ionization. With the exception of vibrational as a function of distance away from the particle. Note that at
excitation, which is not affected by pressure, if the air pressure small distances, the force is strongly repulsive, tending to
is lowered, these onset temperatures decrease; conversely, push the two molecules away from each other. However, as
if the air pressure is increased these onset temperatures are we move further away from the molecule, the intermolecular
raised. force rapidly decreases and becomes a weak attractive force,
tending to attract molecules together. At distances approxi-
mately 10 molecular diameters away from the molecule, the
6.4 Definition of real gases and perfect gases magnitude of the intermolecular force is negligible. Because
the molecules are in constant motion, and this motion is what
Chemically reacting flows are an interesting coupling of gas generates the macroscopic thermodynamic properties of the
dynamics, physical chemistry, statistical thermodynamics, system, the intermolecular force should affect these macro-
kinetic theory, and quantum physics. It is a highly com- scopic properties. This leads to the following definition: Real
plex subject (therefore making it very interesting), and is gas is a gas where intermolecular forces are important and
far too expansive to cover in this chapter. Anderson (2006) must be accounted for.
and Vincenti and Kruger (1965) are reasonable introduc- On the other hand, if the molecules are spaced, on the
tions to the subject. The purpose of this chapter is only to average, more than 10 molecular diameters apart, the mag-
introduce some basic ideas. In particular, we now lay out a nitude of the intermolecular force is very small (see Figure
skeleton of definitions against which the understanding of 12) and can be neglected. This, for example, is the case for
high-temperature flows can begin. air at standard conditions. This leads to the next following
To begin, we distinguish between a real gas and a perfect definition: Perfect gas is a gas where intermolecular forces
gas, defined as follows. Consider the air around you as made are negligible.
up of molecules that are in random motion, frequently col- For most problems in aerodynamics, the assumption of
liding with neighboring molecules. Imagine that you pluck a perfect gas is very reasonable. Conditions that require

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 27

the assumption of a real gas are very high pressures (p ≈ Hence, h and e are functions of T only, that is,
1000atm) and/or low temperatures (T ≈ 30K). Under these
conditions, the molecules in the system will be packed closely h = h(T )
together and will be moving slowly with consequent low (127)
e = e(T )
inertia. Thus, the intermolecular force has every opportunity
to act on the molecules in the system and in turn to mod-
The perfect-gas equation of state holds, for example,
ify the macroscopic properties of the system. In contrast, at
lower pressures (p ≈ 10atm, for example) and higher tem-
peratures (e.g., T = 300K), the molecules are widely spaced pv = RT (128)
apart and are moving more rapidly with consequent higher
inertia. Thus, on the average, the intermolecular force has where R is a constant. The temperature variation of spe-
little effect on the particle motion and therefore on the macro- cific heats, hence the whole nature of a thermally perfect
scopic properties of the system. We can assume such a gas gas, is caused by the excitation of vibrational energy within
to be a perfect gas, where the intermolecular force can be the molecules of the gas and to the electronic energy
ignored. Deviations from perfect-gas behavior tend to be pro- associated with electron motion within the atoms and
portional to p/T 3 , which makes qualitative sense based on molecules.
the preceding discussion.
6.5.3 Chemically reacting mixture of perfect gases
6.5 Classification of gases Here we are dealing with a multispecies, chemically react-
ing gas where intermolecular forces are neglected; hence,
6.5.1 Calorically perfect gas each individual species obeys the perfect-gas equation of
By definition, a calorically perfect gas is one with constant state. At this stage, we need to make a distinction between
specific heats cp and cv . In turn, the ratio of specific heats equilibrium and nonequilibrium chemically reacting gases.
γ = cp /cv is constant. For this gas, the enthalpy and internal Our discussion here will be preliminary; a more fundamental
energy are functions of temperature, given explicitly by understanding of the meaning of equilibrium and nonequilib-
rium systems is discussed in Anderson (2006) and Vincenti
h = cp T (124) and Kruger (1965). For the time being, imagine that you take
the air in the room around you, and instantly increase the
and(116)e = cv T The perfect-gas equation of state holds, for temperature to 5000 K, holding the pressure constant at 1 atm.
example,(117)pv = RT where R is a constant. In the intro- From Figure 11, we know that dissociation will occur. Indeed,
ductory study of compressible flow, the assumption of a let us allow some time (may be several hundred milliseconds)
calorically perfect gas is almost always made. Hence, the for the gas properties to “settle out”, and come to some steady
thermodynamics of a calorically perfect gas is probably quite state at 5000 K and 1 atm. The chemical composition that
familiar to you. finally evolves in the limit of “large” times (milliseconds)
is the equilibrium composition at 5000 K and 1 atm. In con-
trast, during the first few milliseconds immediately after we
6.5.2 Thermally perfect gas instantly increase the temperature to 5000 K, the dissociation
By definition, a thermally perfect gas is one where cp reactions are just beginning to take place, and the variation
and cv are variables and specifically are functions of of the amount of O2 , O, N2 , N, and so on in the gas is chang-
temperature only. ing as a function of time. This is a nonequilibrium system.
After the lapse of a sufficient time, the amounts of O2 , O,
N2 , and so on will approach some steady values, and these
cp = f1 (T )
(125) steady values are the equilibrium values. It is inferred from
cv = f2 (T ) the preceding that once the system is in equilibrium, the equi-
librium values of the mass fractions cO2 , cN2 , cO , cN , and so
Differential changes in the h and e are related to differential on will depend only on the pressure and temperature, that
changes in T via is, at 5000 K and 1 atm, the equilibrium chemical composi-
tion is uniquely defined. In contrast, for the nonequilibrium
dh = cp dT system, cO2 , cN2 , cO , and so on depend not only on p and T
(126)
de = cv dT but also on time. If the nonequilibrium system was a fluid

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
28 Fundamentals of Fluid Flows

element rapidly expanding through a shock-tunnel nozzle, 6.5.4 Real gas


another way of stating this effect is to say that cO2 , cN2 , and
Here, we must take into account the effect of intermolecular
so on depend on the “history” of the flow.
forces. We could formally consider a chemically reacting gas
With these thoughts in mind, we can define a chemi-
as well as a nonreacting real gas. However, in practice, a gas
cally reacting mixture of perfect gases as follows. Consider
behaves as a real gas under conditions of very high pressure
a system at pressure p and temperature T . For convenience,
and low temperature – conditions that accentuate the influ-
assume a unit mass for the system. The number of particles
ence of intermolecular forces on the gas. For these conditions,
of each different chemical species per unit mass of mixture is
the gas is rarely chemically reacting. Therefore, for simplic-
given by N1 , N2 , . . ., Nn . For each individual chemical species
ity, we will consider a nonreacting gas here. Recall that for
present in the mixture (assuming a perfect gas), the enthalpy
both the cases of acalorically perfect gas and a thermally per-
and internal energy per unit mass of i, hi , and ei , respectively,
fect gas, h and e were functions of T only. For a real gas, with
will be functions of T (i.e., each individual species, by itself,
intermolecular forces, h and e depend on p (or v) as well:
behaves as a thermally perfect gas). However, h and e for the
chemically reacting mixture depend not only on hi and ei but
h = h(T, p)
also on how much of each species is present. Therefore, for a
chemically reacting mixture of perfect gases, in the general e = e(T, v)
(132)
nonequilibrlium case, we write cp = f1 (T, p)
cv = f2 (T, v)
h = h(T, N1 , N2 , N3 , . . . , Nn )
e = e(T, N1 , N2 , N3 , . . . , Nn ) Moreover, the perfect-gas equation of state is no longer
(129) valid here. Instead, we must use a real-gas equation of state,
cp = f1 (T, N1 , N2 , N3 , . . . , Nn )
of which there are many versions. Perhaps the most familiar
cv = f2 (T, N1 , N2 , N3 , . . . , Nn ) is the Van der Waals equation, given by
 
where, in general, N1 , N2 , N3 , . . ., Nn depend on p, T , and a
the “history of the gas flow.” The perfect-gas equation of state p+ 2 (v − b) = RT (133)
v
still holds:
a and b are constants that depend on the type of gas. Note
pv = RT (130) that equation (133) reduces to the perfect-gas equation of
state when a = b = 0. In equation (133) the terms a/v2 take
However, here R is a variable because in a chemically reacting into account the intermolecular force effects, and b takes into
gas the molecular weight of the mixture M is a variable, and account the actual volume of the system occupied by the
R = R/M, where R is the universal gas constant. volume of the gas particles themselves.
For the special case of an equilibrium gas, the chemical
composition is a unique function of p and T ; hence, N1 =
6.5.5 Summary
f1 (p, T ), N2 = f2 (p, T ), and so on. Therefore, the preceding
results for h, e, cp , and cv become In summary, the preceding discussion has presented four
different categories of gases. Any existing analyses of ther-
h = h(T, p) modynamic and gas dynamic problems will fall into one of
these categories; they are presented in this chapter so that
e = e1 (T, p) = e2 (T, v)
(131) one can establish an inventory of such gases in mind. It is
cp = f1 (T, p) extremely helpful to keep these categories in mind while
cv = f2 (T, p) = f3 (T, v) performing any study of gas dynamics.
Also, to equate these different categories to a practical
In the preceding, it is frequently convenient to think of e situation, let us once again take the case of air. Imagine
and cv as functions of T and v rather than T and p. It does not that you take the air in the room around you and begin to
make any difference, however, because for a thermodynamic increase its temperature. At room temperature, air is essen-
system in equilibrium (including an equilibrium chemically tially a calorically perfect gas, and it continues to act as a
reacting system) the state of the system is uniquely defined calorically perfect gas until the temperature reaches approx-
by any two state variables. The choice of T and p, or T and imately 800 K. Then, as the temperature increases further,
v, in the preceding, is somewhat arbitrary in this sense. we see from Figure 11 that vibrational excitation becomes

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Governing Equations for Fundamental Aerodynamics 29

important. When this happens, air acts as a thermally perfect ever, in the case of a nonequilibrium chemically reacting
gas. Finally, above 2500 K, chemical reactions occur, and air flow, additional continuity equations are necessary for each
becomes a chemically reacting mixture of perfect gases. If individual chemical species, and these species continuity
we were to go in the opposite direction, that is, reduce the equations involve the finite rate of individual chemical reac-
air temperature considerably below room temperature, and/or tions taking place in the system, and mass diffusion.
increase the pressure to a very high value, say, 1000 atm, then The momentum equation is a statement of Newton’s sec-
the air would behave as a real gas. ond law, which is the same for reacting and nonreacting flows.
Finally, it is important to note a matter of nomenclature. However, the viscosity coefficient in the momentum equation
We have followed classical physical chemistry in defining must be properly calculated for the reacting gas.
a gas where intermolecular forces are important as a real The energy equation is a statement of the first law of
gas. Unfortunately, an ambiguous term has evolved in the thermodynamics, which is the same for reacting and non-
aerodynamic literature that means something quite different. reacting flows. However, again the transport properties of
In the 1950s, aerodynamicists were suddenly confronted with viscosity, thermal conduction, and diffusion coefficient must
hypersonic entry vehicles at velocities as high as 26 000 ft s−1 be properly calculated for the reacting gas. Energy transport
(8 km s−1 ). The shock layers around such vehicles were hot by mass diffusion adds another term in the energy equation,
enough to cause vibrational excitation, dissociation, and even and the value of the internal energy in the energy equa-
ionization. These were “real” effects that happened in air in tion must take into account the changing chemical species
“real life”. Hence, it became fashionable in the aerodynamic with regard to the release or absorption of chemical energy
literature to denote such conditions as real-gas effects. For into the gas. Finally, if the gas temperature is high enough,
example, the categories just itemized as a thermally perfect energy transfer by thermal radiation adds another term to
gas, and as a chemically reacting mixture of perfect gases, the energy equation, and contributes to making the flow
would come under the classification of real-gas effects in nonadiabatic.
some of the aerodynamic literature. But in light of classical
physical chemistry, this is truly a misnomer. A real gas is
truly one in which intermolecular forces are important, and
REFERENCES
this has nothing to do with vibrational excitation or chemical
reactions.
Anderson, John D. Jr (1995) Computational Fluid Dynamics: The
Basics with Applications, McGraw-Hill.
Anderson, John D. Jr (2006) Hypersonic and High-Temperature
6.6 Governing equations for chemically reacting Gas Dynamics, 2nd edn, American Institute of Aeronautics and
flows Astronautics, Reston, VA.
Anderson, (2010) Fundamentals of Aerodynamics, 5th edn,
The governing equations for fluid flows have been developed McGraw-Hill.
in Section 6 for the case of nonreacting flows. How are these Vincenti, W.G. and Kruger, C.H. (1965) Introduction to Physical
equations affected by chemically reacting flows? Here we Gas Dynamics, Wiley, New York.
simply make some overall statements rather than present the
governing equations for chemically reacting flows; the details
are beyond the scope of this chapter. For in-depth discussions
of the equations (see Anderson, 2006; Vincenti and Kruger, FURTHER READING
1965).
The continuity equation is the same for nonreacting and Schlichting, H. (1979) Boundary Layer Theory, 7th edn, McGraw-
reacting flows. Mass is conserved in both situations. How- Hill.

Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002
Encyclopedia of Aerospace Engineering, Online © 2010 John Wiley & Sons, Ltd.
This article is © 2010 John Wiley & Sons, Ltd.
This article was published in the Encyclopedia of Aerospace Engineering in 2010 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470686652.eae002

You might also like