The Curl of a Vector Field. Stokes' Theorem Ex- Presses the Integral Of

Total Page:16

File Type:pdf, Size:1020Kb

The Curl of a Vector Field. Stokes' Theorem Ex- Presses the Integral Of The Curl of a Vector Field. Stokes’ Theorem ex- A related operation, called the divergence will be in- presses the integral of a vector field F around a closed troduced later. curve as a surface integral of another vector field, called the curl of F. This vector field is constructed Each of these acts as a derivative, and there is a version in the proof of the theorem. Once we have it, we in- of the fundamental theorem that evaluates an integral vent the notation ∇×F in order to remember how to of this derivative. compute it. In this notation ∇ stands for the vector The fundamental theorem for line integrals says operator that the difference of a scalar function f (x, y, z) at ∂ ∂ ∂ ∇ , , two points is the integral of f (the gradient of f ) ∂x ∂y ∂z along any path joining the two points. From this it follows that the integral of ∇ f is independent of path ∇ and expressions containing are interpreted by first and that the integral of ∇ f along any closed path is pretending that it is an ordinary vector, so that vector zero. operations will introduce terms that abut a compo- nent from the first factor with one from the second Stokes’ theorem says that the integral of a vector field factor. For numerical vectors, this is interpreted as F(x, y, z) along a closed path is equal to the integral multiplication, but for ∇ the interpretation is to let of ∇×F (the curl of F) along any surface having the the component of ∇ operate on the component from given path as its positively oriented boundary. From the other factor. this it follows that the integral of ∇×F is zero on any closed surface. The gradient also uses this interpretation of ∇ to con- struct a vector field from a scalar function. The curl The divergence theorem (to be discussed later) re- takes one vector field to another using a construction lates the integral of a vector field over a closed surface modeled by the cross product. to the integral of its divergence over the three dimen- 16.5.1 16.5.2 sional region bounded by the surface. Although the #1. hxy, yz, zxi. details are different, all three theorems have similar #5. hex sin y, ex cos y, zi. statements. D E x , y , 1 Special second derivatives. It has already been noted #7. z z z . that For any of these whose curl is zero, express as a gra- ∇×(∇ f ) = 0. dient. Formally, this is just the equality of mixed partials, =∇ Finding a vector field from its curl. The first four but it is tied to Stokes’ Theorem. If F f , the line exercises of Section 16.8 have the form: use Stokes’ integral of F along any curve is the difference of the Theorem to evaluate values of f at the endpoints. For a closed curve, this ZZ is always zero. Stokes’ Theorem then says that the ∇×F dS. surface integral of its curl is zero for every surface, so S it is not surprising that the curl itself is zero. In this form, there isn’t much to the exercise. This Stokes’ theorem also says that the integral of the curl way of stating the exercise gives F, so its curl need of a vector field over a closed surface is zero. If we try never be computed since you must evaluate the line to write a given vector field G as the curl of another integral in order to feel that you have used Stokes’ vector field F, we will meet an obstruction to com- Theorem. There is still something to be done: you pleting the computation in the form of a combination need to produce a parameterization of the positively of derivatives of components of G that must be zero. oriented boundary from a description of the surface, Exercises 16.5. Find curl of the following vector but the statement of the exercise obscures its real con- fields. tent. 16.5.3 16.5.4 It would be more interesting to start from a vector field Looking at the components of the curl, we have G that is of the form G =∇×F, and determine F. Although this is more complicated than the process −Yz = P for writing a conservative vector field F as F =∇f , = (∗) it has a similar flavor. Xz Q − = . We now describe a solution of the equation Yx X y R ∇×F = G If the first of the equations in (∗) is solved for Y (up to a function of x and y), and this result put into the for F when G is given. third equation, we are left with the task of finding It is useful to simplify the problem as much as possible X from its derivatives with respect to y and z. This before beginning. We know that F is only defined up was exactly what we did in recognizing conservative to a term of the form ∇ f , and there is no difficulty (in vector fields as gradients. The only requirement is that principle, although it does require integration) finding the equations Xzy = X yz must hold. These equations a function f for which f3(x, y, z) is any expression. are In particular, any solution F could be replaced by F − Xzy = Q y ∇ f , where f3(x, y, z) is equal to the third component X yz = Yxz − Rz. of F. This means that we need only look for Finally, F = hX, Y, 0i. Yxz = Yzx =−Px , Let so G = hP, Q, Ri. Q y =−Px − Rz. 16.5.5 16.5.6 2 Note that this condition depends only on G and not (since Yx = 0). Then X = xz /2 + f (x, y) and on the particular choice of Y . f2(x, y) =−xy. The solution can be completed be- cause the expression that is supposed to be f (x, y) If this is satisfied, then we can find X for any Y , which 2 really is independent of z. Integrating this, gives completes the determination of F. f (x, y) =−xy2/2 as one solution. Thus, This shows that G = hP, Q, Ri satisfies G =∇×F D E + + = 1 if and only if Px Q y Rz 0. This looks like it F = xz2 − xy2, −yz2, 0 . should be denoted ∇·G. 2 The whole story is a little more subtle. It is easy to If we subtract ∇(x2z2/4) = xz2/2, 0, x2z/2 from verify that ∇×(∇ f ) = 0 and ∇·(∇×F) = 0, but the this, we have the more symmetric solution D E constructive part of these equivalences assumed that 1 the functions obtained in the solution were defined F =− xy2, yz2, zx2 . everywhere. For objects defined only in part of space, 2 the necessary condition my hold but it may not be Additional Exercises. Write the given vector field G possible to perform the construction. as G =∇×F, if possible. An example. From exercise 1 in Section 16.5, we (A) G = h0, 0, 1i. know that∇·G = 0 when G = hyz, xz, xyi. The = 2, 2, first equation of (∗) becomes Yz =−yz, so we take (B) G x y 0 . =− 2/ Y yz 2. This leaves (C) G = h0, 0, xeyi. Xz = xz Why is it called “curl”? Consider the vector field X y =−xy R = h−y, x, 0i 16.5.7 16.5.8 that points in the positive direction along each circle to r 2, so this effect is not overwhelming. However with center at the origin, with a magnitude propor- when considered on a fixed scale, the magnitude of tional to the radius. Such a motion corresponds to our the curl gives a noticeable rotation. To be free of such image of pure rotation. An easy computation shows effects, a vector field must have zero curl. that ∇×R = h0, 0, 2i, identifying the axis of rotation everywhere. A phys- ical model uses a small paddle-wheel moving in the flow R. When the axis of the wheel is lined up with the z-axis, the difference in speeds at the ends of the paddle causes to paddle spin relative to coordinates with fixed directions as its center of mass moves with the flow. This behavior is seen throughout the flow — not just near the center of rotation. The rotation is detected in the vector field R itself by forming the line integral of R along closed curves. Stokes’ theorem tells us that such an integral is given by the integral of ∇×R over a surface bounded by the curve. In particular, the value of the integral may be estimated by the product of the area of the surface and the component of ∇×R perpendicular to the surface. For a small circle of radius r, the area is proportional 16.5.9 16.5.10.
Recommended publications
  • Line, Surface and Volume Integrals
    Line, Surface and Volume Integrals 1 Line integrals Z Z Z Á dr; a ¢ dr; a £ dr (1) C C C (Á is a scalar ¯eld and a is a vector ¯eld) We divide the path C joining the points A and B into N small line elements ¢rp, p = 1;:::;N. If (xp; yp; zp) is any point on the line element ¢rp, then the second type of line integral in Eq. (1) is de¯ned as Z XN a ¢ dr = lim a(xp; yp; zp) ¢ rp N!1 C p=1 where it is assumed that all j¢rpj ! 0 as N ! 1. 2 Evaluating line integrals The ¯rst type of line integral in Eq. (1) can be written as Z Z Z Á dr = i Á(x; y; z) dx + j Á(x; y; z) dy C CZ C +k Á(x; y; z) dz C The three integrals on the RHS are ordinary scalar integrals. The second and third line integrals in Eq. (1) can also be reduced to a set of scalar integrals by writing the vector ¯eld a in terms of its Cartesian components as a = axi + ayj + azk. Thus, Z Z a ¢ dr = (axi + ayj + azk) ¢ (dxi + dyj + dzk) C ZC = (axdx + aydy + azdz) ZC Z Z = axdx + aydy + azdz C C C 3 Some useful properties about line integrals: 1. Reversing the path of integration changes the sign of the integral. That is, Z B Z A a ¢ dr = ¡ a ¢ dr A B 2. If the path of integration is subdivided into smaller segments, then the sum of the separate line integrals along each segment is equal to the line integral along the whole path.
    [Show full text]
  • 1 Curl and Divergence
    Sections 15.5-15.8: Divergence, Curl, Surface Integrals, Stokes' and Divergence Theorems Reeve Garrett 1 Curl and Divergence Definition 1.1 Let F = hf; g; hi be a differentiable vector field defined on a region D of R3. Then, the divergence of F on D is @ @ @ div F := r · F = ; ; · hf; g; hi = f + g + h ; @x @y @z x y z @ @ @ where r = h @x ; @y ; @z i is the del operator. If div F = 0, we say that F is source free. Note that these definitions (divergence and source free) completely agrees with their 2D analogues in 15.4. Theorem 1.2 Suppose that F is a radial vector field, i.e. if r = hx; y; zi, then for some real number p, r hx;y;zi 3−p F = jrjp = (x2+y2+z2)p=2 , then div F = jrjp . Theorem 1.3 Let F = hf; g; hi be a differentiable vector field defined on a region D of R3. Then, the curl of F on D is curl F := r × F = hhy − gz; fz − hx; gx − fyi: If curl F = 0, then we say F is irrotational. Note that gx − fy is the 2D curl as defined in section 15.4. Therefore, if we fix a point (a; b; c), [gx − fy](a; b; c), measures the rotation of F at the point (a; b; c) in the plane z = c. Similarly, [hy − gz](a; b; c) measures the rotation of F in the plane x = a at (a; b; c), and [fz − hx](a; b; c) measures the rotation of F in the plane y = b at the point (a; b; c).
    [Show full text]
  • Chapter 8 Change of Variables, Parametrizations, Surface Integrals
    Chapter 8 Change of Variables, Parametrizations, Surface Integrals x0. The transformation formula In evaluating any integral, if the integral depends on an auxiliary function of the variables involved, it is often a good idea to change variables and try to simplify the integral. The formula which allows one to pass from the original integral to the new one is called the transformation formula (or change of variables formula). It should be noted that certain conditions need to be met before one can achieve this, and we begin by reviewing the one variable situation. Let D be an open interval, say (a; b); in R , and let ' : D! R be a 1-1 , C1 mapping (function) such that '0 6= 0 on D: Put D¤ = '(D): By the hypothesis on '; it's either increasing or decreasing everywhere on D: In the former case D¤ = ('(a);'(b)); and in the latter case, D¤ = ('(b);'(a)): Now suppose we have to evaluate the integral Zb I = f('(u))'0(u) du; a for a nice function f: Now put x = '(u); so that dx = '0(u) du: This change of variable allows us to express the integral as Z'(b) Z I = f(x) dx = sgn('0) f(x) dx; '(a) D¤ where sgn('0) denotes the sign of '0 on D: We then get the transformation formula Z Z f('(u))j'0(u)j du = f(x) dx D D¤ This generalizes to higher dimensions as follows: Theorem Let D be a bounded open set in Rn;' : D! Rn a C1, 1-1 mapping whose Jacobian determinant det(D') is everywhere non-vanishing on D; D¤ = '(D); and f an integrable function on D¤: Then we have the transformation formula Z Z Z Z ¢ ¢ ¢ f('(u))j det D'(u)j du1::: dun = ¢ ¢ ¢ f(x) dx1::: dxn: D D¤ 1 Of course, when n = 1; det D'(u) is simply '0(u); and we recover the old formula.
    [Show full text]
  • Curl, Divergence and Laplacian
    Curl, Divergence and Laplacian What to know: 1. The definition of curl and it two properties, that is, theorem 1, and be able to predict qualitatively how the curl of a vector field behaves from a picture. 2. The definition of divergence and it two properties, that is, if div F~ 6= 0 then F~ can't be written as the curl of another field, and be able to tell a vector field of clearly nonzero,positive or negative divergence from the picture. 3. Know the definition of the Laplace operator 4. Know what kind of objects those operator take as input and what they give as output. The curl operator Let's look at two plots of vector fields: Figure 1: The vector field Figure 2: The vector field h−y; x; 0i: h1; 1; 0i We can observe that the second one looks like it is rotating around the z axis. We'd like to be able to predict this kind of behavior without having to look at a picture. We also promised to find a criterion that checks whether a vector field is conservative in R3. Both of those goals are accomplished using a tool called the curl operator, even though neither of those two properties is exactly obvious from the definition we'll give. Definition 1. Let F~ = hP; Q; Ri be a vector field in R3, where P , Q and R are continuously differentiable. We define the curl operator: @R @Q @P @R @Q @P curl F~ = − ~i + − ~j + − ~k: (1) @y @z @z @x @x @y Remarks: 1.
    [Show full text]
  • An Introduction to Space–Time Exterior Calculus
    mathematics Article An Introduction to Space–Time Exterior Calculus Ivano Colombaro * , Josep Font-Segura and Alfonso Martinez Department of Information and Communication Technologies, Universitat Pompeu Fabra, 08018 Barcelona, Spain; [email protected] (J.F.-S.); [email protected] (A.M.) * Correspondence: [email protected]; Tel.: +34-93-542-1496 Received: 21 May 2019; Accepted: 18 June 2019; Published: 21 June 2019 Abstract: The basic concepts of exterior calculus for space–time multivectors are presented: Interior and exterior products, interior and exterior derivatives, oriented integrals over hypersurfaces, circulation and flux of multivector fields. Two Stokes theorems relating the exterior and interior derivatives with circulation and flux, respectively, are derived. As an application, it is shown how the exterior-calculus space–time formulation of the electromagnetic Maxwell equations and Lorentz force recovers the standard vector-calculus formulations, in both differential and integral forms. Keywords: exterior calculus; exterior algebra; electromagnetism; Maxwell equations; differential forms; tensor calculus 1. Introduction Vector calculus has, since its introduction by J. W. Gibbs [1] and Heaviside, been the tool of choice to represent many physical phenomena. In mechanics, hydrodynamics and electromagnetism, quantities such as forces, velocities and currents are modeled as vector fields in space, while flux, circulation, divergence or curl describe operations on the vector fields themselves. With relativity theory, it was observed that space and time are not independent but just coordinates in space–time [2] (pp. 111–120). Tensors like the Faraday tensor in electromagnetism were quickly adopted as a natural representation of fields in space–time [3] (pp. 135–144). In parallel, mathematicians such as Cartan generalized the fundamental theorems of vector calculus, i.e., Gauss, Green, and Stokes, by means of differential forms [4].
    [Show full text]
  • Surface Integrals
    VECTOR CALCULUS 16.7 Surface Integrals In this section, we will learn about: Integration of different types of surfaces. PARAMETRIC SURFACES Suppose a surface S has a vector equation r(u, v) = x(u, v) i + y(u, v) j + z(u, v) k (u, v) D PARAMETRIC SURFACES •We first assume that the parameter domain D is a rectangle and we divide it into subrectangles Rij with dimensions ∆u and ∆v. •Then, the surface S is divided into corresponding patches Sij. •We evaluate f at a point Pij* in each patch, multiply by the area ∆Sij of the patch, and form the Riemann sum mn * f() Pij S ij ij11 SURFACE INTEGRAL Equation 1 Then, we take the limit as the number of patches increases and define the surface integral of f over the surface S as: mn * f( x , y , z ) dS lim f ( Pij ) S ij mn, S ij11 . Analogues to: The definition of a line integral (Definition 2 in Section 16.2);The definition of a double integral (Definition 5 in Section 15.1) . To evaluate the surface integral in Equation 1, we approximate the patch area ∆Sij by the area of an approximating parallelogram in the tangent plane. SURFACE INTEGRALS In our discussion of surface area in Section 16.6, we made the approximation ∆Sij ≈ |ru x rv| ∆u ∆v where: x y z x y z ruv i j k r i j k u u u v v v are the tangent vectors at a corner of Sij. SURFACE INTEGRALS Formula 2 If the components are continuous and ru and rv are nonzero and nonparallel in the interior of D, it can be shown from Definition 1—even when D is not a rectangle—that: fxyzdS(,,) f ((,))|r uv r r | dA uv SD SURFACE INTEGRALS This should be compared with the formula for a line integral: b fxyzds(,,) f (())|'()|rr t tdt Ca Observe also that: 1dS |rr | dA A ( S ) uv SD SURFACE INTEGRALS Example 1 Compute the surface integral x2 dS , where S is the unit sphere S x2 + y2 + z2 = 1.
    [Show full text]
  • Arxiv:2109.03250V1 [Astro-Ph.EP] 7 Sep 2021 by TESS
    Draft version September 9, 2021 Typeset using LATEX modern style in AASTeX62 Efficient and precise transit light curves for rapidly-rotating, oblate stars Shashank Dholakia,1 Rodrigo Luger,2 and Shishir Dholakia1 1Department of Astronomy, University of California, Berkeley, CA 2Center for Computational Astrophysics, Flatiron Institute, New York, NY Submitted to ApJ ABSTRACT We derive solutions to transit light curves of exoplanets orbiting rapidly-rotating stars. These stars exhibit significant oblateness and gravity darkening, a phenomenon where the poles of the star have a higher temperature and luminosity than the equator. Light curves for exoplanets transiting these stars can exhibit deviations from those of slowly-rotating stars, even displaying significantly asymmetric transits depending on the system’s spin-orbit angle. As such, these phenomena can be used as a protractor to measure the spin-orbit alignment of the system. In this paper, we introduce a novel semi-analytic method for generating model light curves for gravity-darkened and oblate stars with transiting exoplanets. We implement the model within the code package starry and demonstrate several orders of magnitude improvement in speed and precision over existing methods. We test the model on a TESS light curve of WASP-33, whose host star displays rapid rotation (v sin i∗ = 86:4 km/s). We subtract the host’s δ-Scuti pulsations from the light curve, finding an asymmetric transit characteristic of gravity darkening. We find the projected spin orbit angle is consistent with Doppler +19:0 ◦ tomography and constrain the true spin-orbit angle of the system as ' = 108:3−15:4 .
    [Show full text]
  • Vector Calculus and Differential Forms with Applications To
    Vector Calculus and Differential Forms with Applications to Electromagnetism Sean Roberson May 7, 2015 PREFACE This paper is written as a final project for a course in vector analysis, taught at Texas A&M University - San Antonio in the spring of 2015 as an independent study course. Students in mathematics, physics, engineering, and the sciences usually go through a sequence of three calculus courses before go- ing on to differential equations, real analysis, and linear algebra. In the third course, traditionally reserved for multivariable calculus, stu- dents usually learn how to differentiate functions of several variable and integrate over general domains in space. Very rarely, as was my case, will professors have time to cover the important integral theo- rems using vector functions: Green’s Theorem, Stokes’ Theorem, etc. In some universities, such as UCSD and Cornell, honors students are able to take an accelerated calculus sequence using the text Vector Cal- culus, Linear Algebra, and Differential Forms by John Hamal Hubbard and Barbara Burke Hubbard. Here, students learn multivariable cal- culus using linear algebra and real analysis, and then they generalize familiar integral theorems using the language of differential forms. This paper was written over the course of one semester, where the majority of the book was covered. Some details, such as orientation of manifolds, topology, and the foundation of the integral were skipped to save length. The paper should still be readable by a student with at least three semesters of calculus, one course in linear algebra, and one course in real analysis - all at the undergraduate level.
    [Show full text]
  • Surface Integrals
    V9. Surface Integrals Surface integrals are a natural generalization of line integrals: instead of integrating over a curve, we integrate over a surface in 3-space. Such integrals are important in any of the subjects that deal with continuous media (solids, fluids, gases), as well as subjects that deal with force fields, like electromagnetic or gravitational fields. Though most of our work will be spent seeing how surface integrals can be calculated and what they are used for, we first want to indicate briefly how they are defined. The surface integral of the (continuous) function f(x,y,z) over the surface S is denoted by (1) f(x,y,z) dS . ZZS You can think of dS as the area of an infinitesimal piece of the surface S. To define the integral (1), we subdivide the surface S into small pieces having area ∆Si, pick a point (xi,yi,zi) in the i-th piece, and form the Riemann sum (2) f(xi,yi,zi)∆Si . X As the subdivision of S gets finer and finer, the corresponding sums (2) approach a limit which does not depend on the choice of the points or how the surface was subdivided. The surface integral (1) is defined to be this limit. (The surface has to be smooth and not infinite in extent, and the subdivisions have to be made reasonably, otherwise the limit may not exist, or it may not be unique.) 1. The surface integral for flux. The most important type of surface integral is the one which calculates the flux of a vector field across S.
    [Show full text]
  • 10A– Three Generalizations of the Fundamental Theorem of Calculus MATH 22C
    10A– Three Generalizations of the Fundamental Theorem of Calculus MATH 22C 1. Introduction In the next four sections we present applications of the three generalizations of the Fundamental Theorem of Cal- culus (FTC) to three space dimensions (x, y, z) 3,a version associated with each of the three linear operators,2R the Gradient, the Curl and the Divergence. Since much of classical physics is framed in terms of these three gener- alizations of FTC, these operators are often referred to as the three linear first order operators of classical physics. The FTC in one dimension states that the integral of a function over a closed interval [a, b] is equal to its anti- derivative evaluated between the endpoints of the interval: b f 0(x)dx = f(b) f(a). − Za This generalizes to the following three versions of the FTC in two and three dimensions. The first states that the line integral of a gradient vector field F = f along a curve , (in physics the work done by F) is exactlyr equal to the changeC in its potential potential f across the endpoints A, B of : C F Tds= f(B) f(A). (1) · − ZC The second, called Stokes Theorem, says that the flux of the Curl of a vector field F through a two dimensional surface in 3 is the line integral of F around the curve that S R 1 C 2 bounds : S CurlF n dσ = F Tds (2) · · ZZS ZC And the third, called the Divergence Theorem, states that the integral of the Divergence of F over an enclosed vol- ume is equal to the flux of F outward through the two dimensionalV closed surface that bounds : S V DivF dV = F n dσ.
    [Show full text]
  • Approaching Green's Theorem Via Riemann Sums
    APPROACHING GREEN’S THEOREM VIA RIEMANN SUMS JENNIE BUSKIN, PHILIP PROSAPIO, AND SCOTT A. TAYLOR ABSTRACT. We give a proof of Green’s theorem which captures the underlying intuition and which relies only on the mean value theorems for derivatives and integrals and on the change of variables theorem for double integrals. 1. INTRODUCTION The counterpoint of the discrete and continuous has been, perhaps even since Eu- clid, the essence of many mathematical fugues. Despite this, there are fundamental mathematical subjects where their voices are difficult to distinguish. For example, although early Calculus courses make much of the passage from the discrete world of average rate of change and Riemann sums to the continuous (or, more accurately, smooth) world of derivatives and integrals, by the time the student reaches the cen- tral material of vector calculus: scalar fields, vector fields, and their integrals over curves and surfaces, the voice of discrete mathematics has been obscured by the coloratura of continuous mathematics. Our aim in this article is to restore the bal- ance of the voices by showing how Green’s Theorem can be understood from the discrete point of view. Although Green’s Theorem admits many generalizations (the most important un- doubtedly being the Generalized Stokes’ Theorem from the theory of differentiable manifolds), we restrict ourselves to one of its simplest forms: Green’s Theorem. Let S ⊂ R2 be a compact surface bounded by (finitely many) simple closed piecewise C1 curves oriented so that S is on their left. Suppose that M F = is a C1 vector field defined on an open set U containing S.
    [Show full text]
  • External Aerodynamics of the Magnetosphere
    ’ .., NASA TECHNICAL NOTE NASA TN- I- &: I -I LOAN COPY: R AFWL (W KIRTLAND AFI EXTERNAL AERODYNAMICS OF THE MAGNETOSPHERE by John R. Spreiter, A Zbertu Y. A Zksne, and Audrey L. Summers Awes Research Center Moffett Fie@ CuZ$ NATIONAL AERONAUTICS AND SPACE ADMINISTRATION WASHINGTON, D. C. JUNE 1968 i TECH LIBRARY KAFB, NM Illll1lllll111lll lllll lilll llll lllll 1111 Ill 0133359 EXTERNAL AERODYNAMICS OF THE MAGNETOSPHERE By John R. Spreiter, Alberta Y. Alksne, and Audrey L. Summers Ames Research Center Moffett Field, Calif. NATIONAL AERONAUTICS AND SPACE ADMINISTRATION For sale by the Clearinghouse for Federal Scientific and Technical Information Springfield, Virginia 22151 - CFSTl price $3.00 EXTERNAL AERODYNAMICS OF THE MAGNETOSPHERE* By John R. Spreiter, Alberta Y. Alksne, and Audrey L. Summers Ames Research Center SUMMARY A comprehensive survey is given of the continuum fluid theory of the solar wind and its interaction with the Earth's magnetic field, and the rela- tion between the calculated results and those actually measured in space. A unified basis for the entire discussion is provided by the equations of mag- netohydrodynamics, augmented by relations from kinetic theory for certain small-scale details of the flow. While the full complexity of magnetohydrodynamics is required for the formulation of the model and the establishment of the proper conditions to apply at the magnetosphere boundary, it is shown that the magnetic field actually experienced in space is usually sufficiently small that an adequate approximation to the solution can be obtained by first solving the simpler equations of gasdynamics for the flow and then using the results to calculate the deformation of the magnetic field.
    [Show full text]