12.07.2015 Views

DIFFERENTIAL GEOMETRY: A First Course in Curves and Surfaces

DIFFERENTIAL GEOMETRY: A First Course in Curves and Surfaces

DIFFERENTIAL GEOMETRY: A First Course in Curves and Surfaces

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>DIFFERENTIAL</strong><strong>GEOMETRY</strong>:A <strong>First</strong> <strong>Course</strong> <strong>in</strong><strong>Curves</strong> <strong>and</strong> <strong>Surfaces</strong>Prelim<strong>in</strong>ary VersionSummer, 2006Theodore Shifr<strong>in</strong>University of GeorgiaDedicated to the memory of Shi<strong>in</strong>g-Shen Chern,my adviser <strong>and</strong> friendc○2006 Theodore Shifr<strong>in</strong>No portion of this work may be reproduced <strong>in</strong> any form without writtenpermission of the author.


§1. Examples, Arclength Parametrization 3(e) Now consider the twisted cubic <strong>in</strong> R 3 , illustrated <strong>in</strong> Figure 1.3, given byα(t) =(t, t 2 ,t 3 ), t ∈ R.Its projections <strong>in</strong> the xy-, xz-, <strong>and</strong> yz-coord<strong>in</strong>ate planes are, respectively, y = x 2 , z = x 3 ,<strong>and</strong> z 2 = y 3 (the cuspidal cubic).(f) Our next example is a classic called the cycloid: Itisthe trajectory of a dot on a roll<strong>in</strong>gwheel (circle). Consider the illustration <strong>in</strong> Figure 1.4. Assum<strong>in</strong>g the wheel rolls withoutOPtaFigure 1.4slipp<strong>in</strong>g, the distance it travels along the ground is equal to the length of the circular arcsubtended by the angle through which it has turned. That is, if the radius of the circle is a<strong>and</strong> it has turned through angle t, then the po<strong>in</strong>t of contact with the x-axis, Q, isat unitsto the right. The vector from the orig<strong>in</strong> to the po<strong>in</strong>t P can be expressed as the sum of thePOQCPaCt a cos ta s<strong>in</strong> tFigure 1.5three vectors −→ OQ, −→ QC, <strong>and</strong> −→ CP (see Figure 1.5):−→OP = −→ OQ + −→ QC + −→ CP<strong>and</strong> hence the function=(at, 0)+(0,a)+(−a s<strong>in</strong> t, −a cos t),α(t) =(at − a s<strong>in</strong> t, a − a cos t) =a(t − s<strong>in</strong> t, 1 − cos t),t ∈ Rgives a parametrization of the cycloid.(g) A (circular) helix is the screw-like path of a bug as it walks uphill on a right circular cyl<strong>in</strong>derat a constant slope or pitch. If the cyl<strong>in</strong>der has radius a <strong>and</strong> the slope is b/a, wecan imag<strong>in</strong>edraw<strong>in</strong>g a l<strong>in</strong>e of that slope on a piece of paper 2πa units long, <strong>and</strong> then roll<strong>in</strong>g the paperup <strong>in</strong>to a cyl<strong>in</strong>der. The l<strong>in</strong>e gives one revolution of the helix, as we can see <strong>in</strong> Figure 1.6. Ifwe take the axis of the cyl<strong>in</strong>der to be vertical, the projection of the helix <strong>in</strong> the horizontalplane is a circle of radius a, <strong>and</strong> so we obta<strong>in</strong> the parametrization α(t) =(a cos t, a s<strong>in</strong> t, bt).


4 Chapter 1. <strong>Curves</strong>2πb2πaFigure 1.6Brief review of hyperbolic trigonometric functions. Just as the circle x 2 + y 2 = 1 isparametrized by (cos θ, s<strong>in</strong> θ), the portion of the hyperbola x 2 − y 2 =1ly<strong>in</strong>g to the right ofthe y-axis, as shown <strong>in</strong> Figure 1.7, is parametrized by (cosh t, s<strong>in</strong>h t), wherecosh t = et + e −t<strong>and</strong> s<strong>in</strong>h t = et − e −t.22By analogy with circular trigonometry, we set tanh t = s<strong>in</strong>h t1<strong>and</strong> secht =cosh t cosh t . The(cosh t, s<strong>in</strong>h t)Figure 1.7follow<strong>in</strong>g formulas are easy to check:cosh 2 t − s<strong>in</strong>h 2 t =1, tanh 2 t + sech 2 t =1s<strong>in</strong>h ′ (t) =cosh t, cosh ′ (t) =s<strong>in</strong>h t, tanh ′ (t) =sech 2 t, sech ′ (t) =− tanh t secht.


§1. Examples, Arclength Parametrization 5(h) When a uniform <strong>and</strong> flexible cha<strong>in</strong> hangs from two pegs, its weight is uniformly distributedalong its length. The shape it takes is called a catenary. 1 As we ask the reader to check <strong>in</strong>Exercise 9, the catenary is the graph of f(x) =C cosh(x/C), for any constant C>0. ThisFigure 1.8curve will appear numerous times <strong>in</strong> this course.Example 2. One of the more <strong>in</strong>terest<strong>in</strong>g curves that arises “<strong>in</strong> nature” is the tractrix. 2 Thetraditional story is this: A dog is at the end of a 1-unit leash <strong>and</strong> buries a bone at (0, 1) as hisowner beg<strong>in</strong>s to walk down the x-axis, start<strong>in</strong>g at the orig<strong>in</strong>. The dog tries to get back to the bone,so he always pulls the leash taut as he is dragged along the tractrix by his owner. His pull<strong>in</strong>g theleash taut means that the leash will be tangent to the curve. When the master is at (t, 0), let the(0,1)▽(x,y)tθFigure 1.9dog’s position be (x(t),y(t)), <strong>and</strong> let the leash make angle θ(t) with the positive x-axis. Then wehave x(t) =t + cos θ(t), y(t) =s<strong>in</strong> θ(t), sotan θ(t) = dydx = y′ (t)x ′ (t) = cos θ(t)θ′ (t)1 − s<strong>in</strong> θ(t)θ ′ (t) .Therefore, θ ′ (t) =s<strong>in</strong> θ(t). Separat<strong>in</strong>g variables <strong>and</strong> <strong>in</strong>tegrat<strong>in</strong>g, we have ∫ dθ/ s<strong>in</strong> θ = ∫ dt, <strong>and</strong>so t = − ln(csc θ + cot θ) +c for some constant c. S<strong>in</strong>ce θ = π/2 when t =0,wesee that c =0.Now, s<strong>in</strong>ce csc θ + cot θ = 1+cos θ 2 cos 2 (θ/2)== cot(θ/2), we can rewrite this ass<strong>in</strong> θ 2 s<strong>in</strong>(θ/2) cos(θ/2)t =lntan(θ/2). Thus, we can parametrize the tractrix byα(θ) = ( cos θ +lntan(θ/2), s<strong>in</strong> θ ) ,π/2 ≤ θ


6 Chapter 1. <strong>Curves</strong>Alternatively, s<strong>in</strong>ce tan(θ/2) = e t ,wehaves<strong>in</strong> θ =2s<strong>in</strong>(θ/2) cos(θ/2) =2et1+e 2t = 2e t = secht+ e−t cos θ = cos 2 (θ/2) − s<strong>in</strong> 2 (θ/2) = 1 − e2t1+e 2t = e−t − e te t = − tanh t,+ e−t <strong>and</strong> so we can parametrize the tractrix <strong>in</strong>stead byβ(t) = ( t − tanh t, secht), t ≥ 0. ▽The fundamental concept underly<strong>in</strong>g the geometry of curves is the arclength of a parametrizedcurve.Def<strong>in</strong>ition. If α: [a, b] → R 3 is a parametrized curve, then for any a ≤ t ≤ b, wedef<strong>in</strong>e itsarclength from a to t to be s(t) =∫ ta‖α ′ (u)‖du.arclength of its trajectory—is the <strong>in</strong>tegral of its speed.That is, the distance a particle travels—theAn alternative approach is to start with the follow<strong>in</strong>gDef<strong>in</strong>ition. Let α: [a, b] → R 3 be a (cont<strong>in</strong>uous) parametrized curve. Given a partition P ={a = t 0


§1. Examples, Arclength Parametrization 7Now, us<strong>in</strong>g this def<strong>in</strong>ition, we can prove that the distance a particle travels is the <strong>in</strong>tegral ofits speed. We will need to use the result of Exercise A.2.4.Proposition 1.1. Let α: [a, b] → R 3 be a piecewise-C 1 parametrized curve. Thenlength(α) =∫ ba‖α ′ (t)‖dt.Proof. Forany partition P of [a, b], we havek∑k∑∫ til(α, P) = ‖α(t i ) − α(t i−1 )‖ =α ′ (t)dt∥ t i−1∥ ≤so length(α) ≤i=1∫ bai=1k∑i=1‖α ′ (t)‖dt. The same holds on any <strong>in</strong>terval.∫ tit i−1‖α ′ (t)‖dt =∫ ba‖α ′ (t)‖dt,Now, for a ≤ t ≤ b, def<strong>in</strong>e s(t) tobethe arclength of the curve α on the <strong>in</strong>terval [a, t]. Thenfor h>0wehave‖α(t + h) − α(t)‖h≤s(t + h) − s(t)h≤ 1 h∫ t+ht‖α ′ (u)‖du,s<strong>in</strong>ce s(t + h) − s(t) isthe arclength of the curve α on the <strong>in</strong>terval [t, t + h]. (See Exercise 8 for thefirst <strong>in</strong>equality.) NowTherefore, by the squeeze pr<strong>in</strong>ciple,‖α(t + h) − α(t)‖lim= ‖α ′ 1(t)‖ = limh→0 + hh→0 + hs(t + h) − s(t)lim= ‖α ′ (t)‖.h→0 + h∫ t+ht‖α ′ (u)‖du.A similar argument works for h


8 Chapter 1. <strong>Curves</strong>An important observation from a theoretical st<strong>and</strong>po<strong>in</strong>t is that any regular parametrized curvecan be reparametrized by arclength. For if α is regular, the arclength function s(t) =∫ ta‖α ′ (u)‖duis an <strong>in</strong>creas<strong>in</strong>g function (s<strong>in</strong>ce s ′ (t) =‖α ′ (t)‖ > 0 for all t), <strong>and</strong> therefore has an <strong>in</strong>verse functiont = t(s). Then we can consider the parametrizationNote that the cha<strong>in</strong> rule tells us thatβ(s) =α(t(s)).β ′ (s) =α ′ (t(s))t ′ (s) =α ′ (t(s))/s ′ (t(s)) = α ′ (t(s))/‖α ′ (t(s))‖is everywhere a unit vector; <strong>in</strong> other words, β moves with speed 1.EXERCISES 1.1*1. Parametrize the unit circle (less the po<strong>in</strong>t (−1, 0)) by the length t <strong>in</strong>dicated <strong>in</strong> Figure 1.11.(x,y)(−1,0)tFigure 1.11♯ 2.Consider the helix α(t) =(a cos t, a s<strong>in</strong> t, bt). Calculate α ′ (t), ‖α ′ (t)‖, <strong>and</strong> reparametrize α byarclength.3. Let α(t) = ( √3 1cos t + √ 1 12s<strong>in</strong> t, √3 1cos t, √3 cos t − √ 12s<strong>in</strong> t ) .reparametrize α by arclength.Calculate α ′ (t), ‖α ′ (t)‖, <strong>and</strong>*4. Parametrize the graph y = f(x), a ≤ x ≤ b, <strong>and</strong> show that its arclength is given by thetraditional formula∫ b √length = 1+ ( f ′ (x) ) 2 dx.a5. a. Show that the arclength of the catenary α(t) =(t, cosh t) for 0 ≤ t ≤ b is s<strong>in</strong>h b.b. Reparametrize the catenary by arclength. (H<strong>in</strong>t: F<strong>in</strong>d the <strong>in</strong>verse of s<strong>in</strong>h by us<strong>in</strong>g thequadratic formula.)*6. Consider the curve α(t) =(e t ,e −t , √ 2t). Calculate α ′ (t), ‖α ′ (t)‖, <strong>and</strong> reparametrize α byarclength, start<strong>in</strong>g at t =0.


§1. Examples, Arclength Parametrization 97. F<strong>in</strong>d the arclength of the tractrix, given <strong>in</strong> Example 2, start<strong>in</strong>g at (0, 1) <strong>and</strong> proceed<strong>in</strong>g to anarbitrary po<strong>in</strong>t.♯ 8.Let P, Q ∈ R 3 <strong>and</strong> let α: [a, b] → R 3 be any parametrized curve with α(a) =P , α(b) =Q.Let v = Q − P . Prove that length(α) ≥‖v‖, sothat the l<strong>in</strong>e segment from P to Q gives theshortest possible path. (H<strong>in</strong>t: Consider∫ baα ′ (t) · vdt <strong>and</strong> use the Cauchy-Schwarz <strong>in</strong>equalityu · v ≤‖u‖‖v‖. Ofcourse, with the alternative def<strong>in</strong>ition on p. 6, it’s even easier.)9. Consider a uniform cable with density δ hang<strong>in</strong>g <strong>in</strong> equilibrium. As shown <strong>in</strong> Figure 1.12, thetension forces T(x +∆x), −T(x), <strong>and</strong> the weight of the piece of cable ly<strong>in</strong>g over [x, x +∆x]all balance. If the bottom of the cable is at x =0,T 0 is the magnitude of the tension there,T(x+∆x)θ +∆θ−T(x)θ −gδ∆sxx+∆xFigure 1.12<strong>and</strong> the cable is the graph y = f(x), show that f ′′ (x) = gδ √1+fT ′ (x) 2 . (Remember that0tan θ = f ′ (x).) Lett<strong>in</strong>g ∫ C = T 0 /gδ, show that f(x) =C cosh(x/C) +c for some constant c.du(H<strong>in</strong>t: To <strong>in</strong>tegrate √ , make the substitution u = s<strong>in</strong>h v.)1+u 210. As shown <strong>in</strong> Figure 1.13, Freddy Fl<strong>in</strong>tstone wishes to drive his car with square wheels along astrange road. How should you design the road so that his ride is perfectly smooth, i.e., so thatthe center of his wheel travels <strong>in</strong> a horizontal l<strong>in</strong>e? (H<strong>in</strong>ts: Start with a square with verticesOCQPFigure 1.13at (±1, ±1), with center C at the orig<strong>in</strong>. If α(s) =(x(s),y(s)) is an arclength parametrization


10 Chapter 1. <strong>Curves</strong>of the road, start<strong>in</strong>g at (0, −1), consider the vector −→ OC = −→ OP + −→ PQ+ −→ QC, where P = α(s) isthe po<strong>in</strong>t of contact <strong>and</strong> Q is the midpo<strong>in</strong>t of the edge of the square. Use −→ QP = sα ′ (s) <strong>and</strong> thefact that −→ QC is a unit vector orthogonal to −→ QP . Express the fact that C moves horizontallyto show that s = − y′ (s)x ′ ;you will need to differentiate unexpectedly. Now use the result of(s)Exercise 4 to f<strong>in</strong>d y = f(x). Also see the h<strong>in</strong>t for Exercise 9.)⎧⎨( )t, t s<strong>in</strong>(π/t) , t ≠011. Show that the curve α(t) =has <strong>in</strong>f<strong>in</strong>ite length on [0, 1]. (H<strong>in</strong>t: Considerl(α, P) with P = {0, 1/N, 2/(2N − 1), 1/(N − 1),...,1/2, 2/3,⎩(0, 0), t =01}.)12. Prove that no four dist<strong>in</strong>ct po<strong>in</strong>ts on the twisted cubic (see Example 1(e)) lie on a plane.13. (a special case of a recent American Mathematical Monthly problem) Suppose α: [a, b] → R 2is a smooth parametrized plane curve (perhaps not arclength-parametrized). Prove that if thechord length ‖α(s) − α(t)‖ depends only on |s − t|, then α must be a (subset of) a l<strong>in</strong>e or acircle. (How many derivatives of α do you need to use?)2. Local Theory: Frenet FrameWhat dist<strong>in</strong>guishes a circle or a helix from a l<strong>in</strong>e is their curvature, i.e., the tendency of thecurve to change direction. We shall now see that we can associate to each smooth (C 3 ) arclengthparametrizedcurve α a natural “mov<strong>in</strong>g frame” (an orthonormal basis for R 3 chosen at each po<strong>in</strong>ton the curve, adapted to the geometry of the curve as much as possible).We beg<strong>in</strong> with a fact from vector calculus which will appear throughout this course.Lemma 2.1. Suppose f, g :(a, b) → R 3 are differentiable <strong>and</strong> satisfy f(t) · g(t) =const for allt. Then f ′ (t) · g(t) =−f(t) · g ′ (t). Inparticular,‖f(t)‖ = const if <strong>and</strong> only if f(t) · f ′ (t) =0 for all t.Proof. S<strong>in</strong>ce a function is constant on an <strong>in</strong>terval if <strong>and</strong> only if its derivative is everywherezero, we deduce from the product rule,(f · g) ′ (t) =f ′ (t) · g(t)+f(t) · g ′ (t),that if f · g is constant, then f · g ′ = −f ′ · g. Inparticular, ‖f‖ is constant if <strong>and</strong> only if ‖f‖ 2 = f · fis constant, <strong>and</strong> this occurs if <strong>and</strong> only if f · f ′ =0. □Remark. This result is <strong>in</strong>tuitively clear. If a particle moves on a sphere centered at the orig<strong>in</strong>,then its velocity vector must be orthogonal to its position vector; any component <strong>in</strong> the directionof the position vector would move the particle off the sphere. Similarly, suppose f <strong>and</strong> g haveconstant length <strong>and</strong> a constant angle between them. Then <strong>in</strong> order to ma<strong>in</strong>ta<strong>in</strong> the constant angle,as f turns towards g, wesee that g must turn away from f at the same rate.Us<strong>in</strong>g Lemma 2.1 repeatedly, we now construct the Frenet frame of suitable regular curves. Weassume throughout that the curve α is parametrized by arclength. Then, for starters, α ′ (s) isthe


§2. Local Theory: Frenet Frame 11unit tangent vector to the curve, which we denote by T(s). S<strong>in</strong>ce T has constant length, T ′ (s) will beorthogonal to T(s). Assum<strong>in</strong>g T ′ (s) ≠ 0, def<strong>in</strong>e the pr<strong>in</strong>cipal normal vector N(s) =T ′ (s)/‖T ′ (s)‖<strong>and</strong> the curvature κ(s) =‖T ′ (s)‖. Sofar, we haveT ′ (s) =κ(s)N(s).If κ(s) =0,the pr<strong>in</strong>cipal normal vector is not def<strong>in</strong>ed. Assum<strong>in</strong>g κ ≠0,wecont<strong>in</strong>ue. Def<strong>in</strong>e theb<strong>in</strong>ormal vector B(s) =T(s) × N(s). Then {T(s), N(s), B(s)} form a right-h<strong>and</strong>ed orthonormalbasis for R 3 .Now, N ′ (s) must be a l<strong>in</strong>ear comb<strong>in</strong>ation of T(s), N(s), <strong>and</strong> B(s). But we know from Lemma2.1 that N ′ (s) · N(s) = 0 <strong>and</strong> N ′ (s) · T(s) = −T ′ (s) · N(s) = −κ(s). We def<strong>in</strong>e the torsionτ(s) =N ′ (s) · B(s). This gives usN ′ (s) =−κ(s)T(s)+τ(s)B(s).F<strong>in</strong>ally, B ′ (s) must be a l<strong>in</strong>ear comb<strong>in</strong>ation of T(s), N(s), <strong>and</strong> B(s). Lemma 2.1 tells us thatB ′ (s) · B(s) =0,B ′ (s) · T(s) =−T ′ (s) · B(s) =0,<strong>and</strong> B ′ (s) · N(s) =−N ′ (s) · B(s) =−τ(s). Thus,In summary, we have:B ′ (s) =−τ(s)N(s).Frenet formulasT ′ (s) =N ′ (s) =−κ(s)T(s)B ′ (s) =κ(s)N(s)+ τ(s)B(s)−τ(s)N(s)The skew-symmetry of these equations is made clearest when we state the Frenet formulas <strong>in</strong>matrix form: ⎡ ⎤ ⎡⎤ ⎡⎤| | | | | | 0 −κ(s) 0⎢ ⎣ T ′ (s) N ′ (s) B ′ ⎥ ⎢⎥ ⎢⎥(s) ⎦ = ⎣ T(s) N(s) B(s) ⎦ ⎣ κ(s) 0 −τ(s) ⎦ .| | | | | | 0 τ(s) 0Indeed, note that the coefficient matrix appear<strong>in</strong>g on the right is skew-symmetric. This is the casewhenever we differentiate an orthogonal matrix depend<strong>in</strong>g on a parameter (s <strong>in</strong> this case). (SeeExercise A.1.4.)Note that, by def<strong>in</strong>ition, the curvature, κ, isalways nonnegative; the torsion, τ, however, has asign, as we shall now see.Example 1. Consider the helix, given by its arclength parametrization (see Exercise 1.1.2)α(s) = ( a cos(s/c),as<strong>in</strong>(s/c), bs/c ) , where c = √ a 2 + b 2 . Then we haveT(s) = 1 ( )−a s<strong>in</strong>(s/c),acos(s/c),bcT ′ (s) = 1 ( ) a ( )−a cos(s/c), −a s<strong>in</strong>(s/c), 0 = − cos(s/c), − s<strong>in</strong>(s/c), 0c 2 }{{} c 2.} {{ }κN


12 Chapter 1. <strong>Curves</strong>Summariz<strong>in</strong>g,κ(s) = a c 2 =Now we deal with B <strong>and</strong> the torsion:aa 2 + b 2 <strong>and</strong> N(s) = ( − cos(s/c), − s<strong>in</strong>(s/c), 0 ) .B(s) =T(s) × N(s) = 1 c(b s<strong>in</strong>(s/c), −b cos(s/c),a)B ′ (s) = 1 c 2 (b cos(s/c),bs<strong>in</strong>(s/c), 0)= −bc 2 N(s),so we <strong>in</strong>fer that τ(s) = bc 2 = ba 2 + b 2 .Note that both the curvature <strong>and</strong> the torsion are constants. The torsion is positive when thehelix is “right-h<strong>and</strong>ed” (b >0) <strong>and</strong> negative when the helix is “left-h<strong>and</strong>ed” (b


§2. Local Theory: Frenet Frame 13Us<strong>in</strong>g the more casual—but convenient—Leibniz notation for derivatives,dTdt = dTdTdsdT= υκN or κN =ds dt ds = dt= 1 ds υExample 2. Let’s calculate the curvature of the tractrix (see Example 2 <strong>in</strong> Section 1). Us<strong>in</strong>gthe first parametrization, we have α ′ (θ) =(− s<strong>in</strong> θ + csc θ, cos θ), <strong>and</strong> sodtdTdt .υ(θ) =‖α ′ (θ)‖ = √ (− s<strong>in</strong> θ + csc θ) 2 + cos 2 θ = √ csc 2 θ − 1=− cot θ.(Note the negative sign because π 2≤ θ0 <strong>and</strong> (s<strong>in</strong> θ, − cos θ) isaunit vector we conclude thatκ(θ) =− tan θ <strong>and</strong> N(θ) =(s<strong>in</strong> θ, − cos θ).Later on we will see an <strong>in</strong>terest<strong>in</strong>g geometric consequence of the equality of the curvature <strong>and</strong> the(absolute value of) the slope. ▽Example 3. Let’s calculate the “Frenet apparatus” for the parametrized curveα(t) =(3t − t 3 , 3t 2 , 3t + t 3 ).We beg<strong>in</strong> by calculat<strong>in</strong>g α ′ <strong>and</strong> determ<strong>in</strong><strong>in</strong>g the unit tangent vector T <strong>and</strong> speed υ:Nowα ′ (t) =3(1 − t 2 , 2t, 1+t 2 ),υ(t) =‖α ′ (t)‖ =3 √ (1 − t 2 ) 2 +(2t) 2 +(1+t 2 ) 2 =3 √ 2(1 + t 2 ) 2 =3 √ 2(1 + t 2 )T(t) = √ 1 12 1+t 2 (1 − t2 , 2t, 1+t 2 )= √ 1 ( )1 − t22 1+t 2 , 2t1+t 2 , 1 .κN = dT dTds = dtdsdtso= 1 dTυ(t) dt(1 −4t√2 (1 + t 2 ) 2 , 2(1 − )t2 )(1 + t 2 ) 2 , 01=3 √ 2(1 + t 2 )1=3 √ 1 2√ ·2(1 + t 2 ) 2 1+t 2} {{ }κ(−2t1+t 2 , 1 − )t21+t 2 , 0 .} {{ }NHere we have factored out the length of the derivative vector <strong>and</strong> left ourselves with a unit vector<strong>in</strong> its direction, which must be the pr<strong>in</strong>cipal normal N; the magnitude that is left must be the<strong>and</strong>


14 Chapter 1. <strong>Curves</strong>curvature κ. Insummary, so far we haveκ(t) =13(1 + t 2 ) 2 <strong>and</strong> N(t) =(−2t1+t 2 , 1 − )t21+t 2 , 0 .Next we f<strong>in</strong>d the b<strong>in</strong>ormal B by calculat<strong>in</strong>g the cross productB(t) =T(t) × N(t) = √ 1 (− 1 − t22 1+t 2 , − 2t )1+t 2 , 1 .And now, at long last, we calculate the torsion by differentiat<strong>in</strong>g B:so τ(t) =κ(t) =−τN = dB dBds = dtds13(1 + t 2 ) 2 . ▽dt= 1 dBυ(t) dt()1 4t√2 (1 + t 2 ) 2 , 2(t2 − 1)(1 + t 2 ) 2 , 01=3 √ 2(1 + t 2 )(1= −3(1 + t 2 )} {{ 2 −2t1+t}2 , 1 − )t21+t 2 , 0 ,} {{ }τNNow we see that curvature enters naturally when we compute the acceleration of a mov<strong>in</strong>gparticle. Differentiat<strong>in</strong>g the formula (∗) onp.12, we obta<strong>in</strong>α ′′ (t) =υ ′ (t)T(s(t)) + υ(t)T ′ (s(t))s ′ (t)= υ ′ (t)T(s(t)) + υ(t) 2( κ(s(t))N(s(t)) ) .Suppress<strong>in</strong>g the variables for a moment, we can rewrite this equation as(∗∗)α ′′ = υ ′ T + κυ 2 N.The tangential component of acceleration is the derivative of speed; the normal component (the“centripetal acceleration” <strong>in</strong> the case of circular motion) is the product of the curvature of the path<strong>and</strong> the square of the speed. Thus, from the physics of the motion we can recover the curvature ofthe path:Proposition 2.2. Forany regular parametrized curve α, wehave κ = ‖α′ × α ′′ ‖‖α ′ ‖ 3 .Proof. S<strong>in</strong>ce α ′ × α ′′ =(υT) × (υ ′ T + κυ 2 N)=κυ 3 T × N <strong>and</strong> κυ 3 > 0, we obta<strong>in</strong> κυ 3 =‖α ′ × α ′′ ‖, <strong>and</strong> so κ = ‖α ′ × α ′′ ‖/υ 3 ,asdesired. □We next proceed to study various theoretical consequences of the Frenet formulas.Proposition 2.3. A space curve is a l<strong>in</strong>e if <strong>and</strong> only if its curvature is everywhere 0.Proof. The general l<strong>in</strong>e is given by α(s) =sv + c for some unit vector v <strong>and</strong> constant vectorc. Then α ′ (s) =T(s) =v is constant, so κ =0. Conversely, if κ =0,then T(s) =T 0 is a constant∫ svector, <strong>and</strong>, <strong>in</strong>tegrat<strong>in</strong>g, we obta<strong>in</strong> α(s) = T(u)du = sT 0 + c for some constant vector c. This0is, once aga<strong>in</strong>, the parametric equation of a l<strong>in</strong>e. □


§2. Local Theory: Frenet Frame 15Example 4. Suppose all the tangent l<strong>in</strong>es of a space curve pass through a fixed po<strong>in</strong>t. Whatcan we say about the curve? Without loss of generality, we take the fixed po<strong>in</strong>t to be the orig<strong>in</strong><strong>and</strong> the curve to be arclength-parametrized by α. Then for every s we have α(s) =λ(s)T(s) forsome scalar function λ. Differentiat<strong>in</strong>g, we haveT(s) =α ′ (s) =λ ′ (s)T(s)+λ(s)T ′ (s) =λ ′ (s)T(s)+λ(s)κ(s)N(s).Then (λ ′ (s) − 1)T(s) +λ(s)κ(s)N(s) =0, so, s<strong>in</strong>ce T(s) <strong>and</strong> N(s) are l<strong>in</strong>early <strong>in</strong>dependent, we<strong>in</strong>fer that λ(s) =s + c for some constant c <strong>and</strong> κ(s) =0. Therefore, the curve must be a l<strong>in</strong>ethrough the fixed po<strong>in</strong>t. ▽Somewhat more challeng<strong>in</strong>g is the follow<strong>in</strong>gProposition 2.4. A space curve is planar if <strong>and</strong> only if its torsion is everywhere 0. The onlyplanar curves with nonzero constant curvature are (portions of) circles.Proof. If a curve lies <strong>in</strong> a plane P, then T(s) <strong>and</strong> N(s) span the plane P 0 parallel to P<strong>and</strong> pass<strong>in</strong>g through the orig<strong>in</strong>. Therefore, B = T × N is a constant vector (the normal toP 0 ), <strong>and</strong> so B ′ = −τN = 0, from which we conclude that τ =0. Conversely, if τ =0,theb<strong>in</strong>ormal vector B is a constant vector B 0 .Now, consider the function f(s) =α(s) · B 0 ;wehavef ′ (s) =α ′ (s) · B 0 = T(s) · B(s) =0,<strong>and</strong> so f(s) =c for some constant c. This means that α lies<strong>in</strong> the plane x · B 0 = c.We leave it to the reader to check <strong>in</strong> Exercise 2a. that a circle of radius a has constant curvature1/a. (This can also be deduced as a special case of the calculation <strong>in</strong> Example 1.) Now suppose aplanar curve α has constant curvature κ 0 . Consider the auxiliary function β(s) =α(s)+ 1 N(s).κ 0Then we have β ′ (s) =α ′ (s) + 1 (−κ 0 (s)T(s)) = T(s) − T(s) =0. Therefore β is a constantκ 0function, say β(s) =P for all s. Now we claim that α is a (subset of a) circle centered at P , for‖α(s) − P ‖ = ‖α(s) − β(s)‖ =1/κ 0 . □We have already seen that a circular helix has constant curvature <strong>and</strong> torsion. We leave itto the reader to check <strong>in</strong> Exercise 10 that these are the only curves with constant curvature <strong>and</strong>torsion. Somewhat more <strong>in</strong>terest<strong>in</strong>g are the curves for which τ/κ is a constant.A generalized helix is a space curve with κ ≠0all of whose tangent vectors make a constantangle with a fixed direction. As shown <strong>in</strong> Figure 2.2, this curve lies on a generalized cyl<strong>in</strong>der,formed by tak<strong>in</strong>g the union of the l<strong>in</strong>es (rul<strong>in</strong>gs) <strong>in</strong> that fixed direction through each po<strong>in</strong>t of thecurve. We can now characterize generalized helices by the follow<strong>in</strong>gProposition 2.5. A curve is a generalized helix if <strong>and</strong> only if τ/κ is constant.Proof. Suppose α is an arclength-parametrized generalized helix. Then there is a (constant)unit vector A with the property that T · A = cos θ for some constant θ. Differentiat<strong>in</strong>g, we obta<strong>in</strong>κN · A =0,whence N · A =0. Differentiat<strong>in</strong>g yet aga<strong>in</strong>, we have(†)(−κT + τB) · A =0.Now, note that A lies <strong>in</strong> the plane spanned by T <strong>and</strong> B, <strong>and</strong> thus B · A = ± s<strong>in</strong> θ. Thus, we <strong>in</strong>ferfrom equation (†) that τ/κ = ± cot θ, which is <strong>in</strong>deed constant.


16 Chapter 1. <strong>Curves</strong>Figure 2.2Conversely, if τ/κ is constant, set τ/κ = cot θ for some angle θ ∈ (0,π). Set A(s) =cos θT(s)+s<strong>in</strong> θB(s). Then A ′ (s) =(κ cos θ − τ s<strong>in</strong> θ)N(s) =0, soA(s) isaconstant unit vector A, <strong>and</strong>T(s) · A = cos θ is constant, as desired. □Example 5. In Example 3 we saw a curve α with κ = τ, sofrom the proof of Proposition 2.5 wesee that the curve should make a constant angle θ = π/4 with the vector A = 1 √2(T+B) =(0, 0, 1)(as should have been obvious from the formula for T alone). We verify this <strong>in</strong> Figure 2.3 by draw<strong>in</strong>gα along with the vertical cyl<strong>in</strong>der built on the projection of α onto the xy-plane. ▽Figure 2.3


§2. Local Theory: Frenet Frame 17The Frenet formulas actually characterize the local picture of a space curve.Proposition 2.6 (Local canonical form). Let α beasmooth (C 3 or better) arclength-parametrizedcurve. If α(0) = 0, then for s near 0, wehave() ( )α(s) = s − κ2 0κ06 s3 + ... T(0) +2 s2 + κ′ (0κ0 τ)06 s3 + ... N(0) +6 s3 + ... B(0).(Here κ 0 , τ 0 , <strong>and</strong> κ ′ 0 denote, respectively, the values of κ, τ, <strong>and</strong> κ′ at 0, <strong>and</strong> lims→0.../s 3 =0.)Proof. Us<strong>in</strong>g Taylor’s Theorem, we writeα(s) =α(0) + sα ′ (0) + 1 2 s2 α ′′ (0) + 1 6 s3 α ′′′ (0) + ...,where lims→0.../s 3 =0. Now,α(0) = 0, α ′ (0) = T(0), <strong>and</strong> α ′′ (0) = T ′ (0) = κ 0 N(0). Differentiat<strong>in</strong>gaga<strong>in</strong>, we have α ′′′ (0) = (κN) ′ (0) = κ ′ 0 N(0) + κ 0(−κ 0 T(0) + τ 0 B(0)). Substitut<strong>in</strong>g, we obta<strong>in</strong>as required.α(s) =sT(0) + 1 2 s2 κ 0 N(0) + 1 ( 6 s3 −κ 2 0T(0) + κ ′ 0N(0) + κ 0 τ 0 B(0) ) + ...() ( )= s − κ2 0κ06 s3 + ... T(0) +2 s2 + κ′ (0κ0 τ)06 s3 + ... N(0) +6 s3 + ... B(0),□We now <strong>in</strong>troduce three fundamental planes at P = α(0):(i) the osculat<strong>in</strong>g plane, spanned by T(0) <strong>and</strong> N(0),(ii) the rectify<strong>in</strong>g plane, spanned by T(0) <strong>and</strong> B(0), <strong>and</strong>(iii) the normal plane, spanned by N(0) <strong>and</strong> B(0).We see that, locally, the projections of α <strong>in</strong>to these respective planes look like(i) (u, (κ 0 /2)u 2 + ...)(ii) (u, (κ 0 τ 0 /6)u 3 + ...), <strong>and</strong>(iii) (u 2 ,( √2τ03 √ κ 0)u 3 + ...),where limu→0.../u 3 =0.Thus, the projections of α <strong>in</strong>to these planes look locally as shown <strong>in</strong> Figure2.4. The osculat<strong>in</strong>g (“kiss<strong>in</strong>g”) plane is the plane that comes closest to conta<strong>in</strong><strong>in</strong>g α near P (seeNBBTTNosculat<strong>in</strong>g plane rectify<strong>in</strong>g plane normal planeFigure 2.4


18 Chapter 1. <strong>Curves</strong>also Exercise 23); the rectify<strong>in</strong>g (“straighten<strong>in</strong>g”) plane is the one that comes closest to flatten<strong>in</strong>gthe curve near P ; the normal plane is normal (perpendicular) to the curve at P . (Cf. Figure 1.3.)EXERCISES 1.21. Compute the curvature of the follow<strong>in</strong>g arclength-parametrized curves:a. α(s) = ( √2 1 1cos s, √2 cos s, s<strong>in</strong> s )b. α(s) = (√ 1+s 2 , ln(s + √ 1+s 2 ) )*c. α(s) = ( 13 (1 + s)3/2 , 1 3 (1 − s)3/2 1, √2 s ) , s ∈ (−1, 1)2. Calculate the unit tangent vector, pr<strong>in</strong>cipal normal, <strong>and</strong> curvature of the follow<strong>in</strong>g curves:a. a circle of radius a: α(t) =(a cos t, a s<strong>in</strong> t)b. α(t) =(t, cosh t)c. α(t) =(cos 3 t, s<strong>in</strong> 3 t), t ∈ (0,π/2)3. Calculate the Frenet apparatus (T, κ, N, B, <strong>and</strong> τ) ofthe follow<strong>in</strong>g curves:*a. α(s) = ( 13 (1 + s)3/2 , 1 3 (1 − s)3/2 1, √2 s ) , s ∈ (−1, 1)b. α(t) = ( 12 et (s<strong>in</strong> t + cos t), 1 2 et (s<strong>in</strong> t − cos t),e t)*c. α(t) = (√ 1+t 2 ,t,ln(t + √ 1+t 2 ) )d. α(t) =(e t cos t, e t s<strong>in</strong> t, e t )e. α(t) =(cosh t, s<strong>in</strong>h t, t)♯ 4. Prove that the curvature of the plane curve y = f(x) isgiven by κ =♯ *5.|f ′′ |(1 + f ′2 ) 3/2 .Use Proposition 2.2 <strong>and</strong> the second parametrization of the tractrix given <strong>in</strong> Example 2 ofSection 1 to recompute the curvature.*6. By differentiat<strong>in</strong>g the equation B = T × N, derive the equation B ′ = −τN.♯ 7. Suppose α is an arclength-parametrized space curve with the property that ‖α(s)‖ ≤‖α(s 0 )‖ =R for all s sufficiently close to s 0 . Prove that κ(s 0 ) ≥ 1/R. (H<strong>in</strong>t: Consider the functionf(s) =‖α(s)‖ 2 . What do you know about f ′′ (s 0 )?)8. Let α be a regular (arclength-parametrized) curve with nonzero curvature. The normal l<strong>in</strong>e toα at α(s) isthe l<strong>in</strong>e through α(s) with direction vector N(s). Suppose all the normal l<strong>in</strong>es toα pass through a fixed po<strong>in</strong>t. What can you say about the curve?9. a. Prove that if all the normal planes of a curve pass through a particular po<strong>in</strong>t, then thecurve lies on a sphere. (H<strong>in</strong>t: Apply Lemma 2.1.)*b. Prove that if all the osculat<strong>in</strong>g planes of a curve pass through a particular po<strong>in</strong>t, then thecurve is planar.10. Prove that if κ = κ 0 <strong>and</strong> τ = τ 0 are nonzero constants, then the curve is a (right) circular helix.(H<strong>in</strong>t: The only solutions of the differential equation y ′′ +k 2 y =0are y = c 1 cos(kt)+c 2 s<strong>in</strong>(kt).)


§2. Local Theory: Frenet Frame 19Remark. It is an amus<strong>in</strong>g exercise to give a <strong>and</strong> b (<strong>in</strong> our formula for the circular helix)<strong>in</strong> terms of κ 0 <strong>and</strong> τ 0 .*11. Proceed as <strong>in</strong> the derivation of Proposition 2.2 to show thatτ = α′ · (α ′′ × α ′′′ )‖α ′ × α ′′ ‖ 2 .12. Let α be a C 4 arclength-parametrized curve with κ ≠0. Prove that α is a generalized helix if<strong>and</strong> only if α ′′ · (α ′′′ × α (iv) )=0. (Here α (iv) denotes the fourth derivative of α.)13. Suppose κτ ≠0atP .Ofall the planes conta<strong>in</strong><strong>in</strong>g the tangent l<strong>in</strong>e to α at P , show that α lieslocally on both sides only of the osculat<strong>in</strong>g plane.14. Let α be a regular curve with κ ≠0atP . Prove that the planar curve obta<strong>in</strong>ed by project<strong>in</strong>gα <strong>in</strong>to its osculat<strong>in</strong>g plane at P has the same curvature at P as α.15. A closed, planar curve C is said to have constant breadth µ if the distance between paralleltangent l<strong>in</strong>es to C is always µ. (No, C needn’t be a circle. See Figure 2.5.) Assume for therest of this problem that the curve is C 2 <strong>and</strong> κ ≠0.(the Wankel eng<strong>in</strong>e design)Figure 2.5a. Let’s call two po<strong>in</strong>ts with parallel tangent l<strong>in</strong>es opposite. Prove that if C has constantbreadth µ, then the chord jo<strong>in</strong><strong>in</strong>g opposite po<strong>in</strong>ts is normal to the curve at both po<strong>in</strong>ts.(H<strong>in</strong>t: If β(s) isopposite α(s), then β(s) =α(s)+λ(s)T(s)+µN(s). <strong>First</strong> expla<strong>in</strong> whythe coefficient of N is µ; then show that λ = 0.)b. Prove that the sum of the reciprocals of the curvature at opposite po<strong>in</strong>ts is equal to µ.(Warn<strong>in</strong>g: If α is arclength-parametrized, β is quite unlikely to be.)16. Let α <strong>and</strong> β be two regular curves def<strong>in</strong>ed on [a, b]. We say β is an <strong>in</strong>volute of α if, for eacht ∈ [a, b],(i) β(t) lies on the tangent l<strong>in</strong>e to α at α(t), <strong>and</strong>(ii) the tangent vectors to α <strong>and</strong> β at α(t) <strong>and</strong> β(t), respectively, are perpendicular.Reciprocally, we also refer to α as an evolute of β.a. Suppose α is arclength-parametrized. Show that β is an <strong>in</strong>volute of α if <strong>and</strong> only ifβ(s) =α(s)+(c − s)T(s) for some constant c (here T(s) =α ′ (s)). We will normally refer


20 Chapter 1. <strong>Curves</strong>to the curve β obta<strong>in</strong>ed with c =0asthe <strong>in</strong>volute of α. Ifyou were to wrap a str<strong>in</strong>garound the curve α, start<strong>in</strong>g at s =0,the <strong>in</strong>volute is the path the end of the str<strong>in</strong>g followsas you unwrap it, always pull<strong>in</strong>g the str<strong>in</strong>g taut, as illustrated <strong>in</strong> the case of a circle <strong>in</strong>Figure 2.6.PθFigure 2.6b. Show that the <strong>in</strong>volute of a helix is a plane curve.c. Show that the <strong>in</strong>volute of a catenary is a tractrix. (H<strong>in</strong>t: You do not need an arclengthparametrization!)d. If α is an arclength-parametrized plane curve, prove that the curve β given byβ(s) =α(s)+ 1κ(s) N(s)is the unique evolute of α ly<strong>in</strong>g <strong>in</strong> the plane of α. Prove, moreover, that this curve isregular if κ ′ ≠0. (H<strong>in</strong>t: Go back to the orig<strong>in</strong>al def<strong>in</strong>ition.)17. F<strong>in</strong>d the <strong>in</strong>volute of the cycloid α(t) =(t + s<strong>in</strong> t, 1 − cos t), t ∈ [−π, π], us<strong>in</strong>g t =0asyourstart<strong>in</strong>g po<strong>in</strong>t. Give a geometric description of your answer.18. Let α be a curve parametrized by arclength with κ, τ ≠0.a. Suppose α lies on the surface of a sphere centered at the orig<strong>in</strong> (i.e., ‖α(s)‖ = const forall s). Prove that(⋆)( ( )τ 1 1 ′ ) ′κ + =0.τ κ(H<strong>in</strong>t: Write α = λT + µN + νB for some functions λ, µ, <strong>and</strong> ν, differentiate, <strong>and</strong> use thefact that {T, N, B} is a basis for R 3 .)b. Prove the converse: If α satisfies the differential equation (⋆), then α lies on the surfaceof some sphere. (H<strong>in</strong>t: Us<strong>in</strong>g the values of λ, µ, <strong>and</strong> ν you obta<strong>in</strong>ed <strong>in</strong> part a, show thatα − (λT + µN + νB) isaconstant vector, the c<strong>and</strong>idate for the center of the sphere.)19. Two dist<strong>in</strong>ct parametrized curves α <strong>and</strong> β are called Bertr<strong>and</strong> mates if for each t, the normall<strong>in</strong>e to α at α(t) equals the normal l<strong>in</strong>e to β at β(t). An example is pictured <strong>in</strong> Figure 2.7.Suppose α <strong>and</strong> β are Bertr<strong>and</strong> mates.


§2. Local Theory: Frenet Frame 21Figure 2.7a. If α is arclength-parametrized, show that β(s) =α(s)+r(s)N(s) <strong>and</strong> r(s) =const. Thus,correspond<strong>in</strong>g po<strong>in</strong>ts of α <strong>and</strong> β are a constant distance apart.b. Show that, moreover, the angle between the tangent vectors to α <strong>and</strong> β at correspond<strong>in</strong>gpo<strong>in</strong>ts is constant. (H<strong>in</strong>t: If T α <strong>and</strong> T β are the unit tangent vectors to α <strong>and</strong> βrespectively, consider T α · T β .)c. Suppose α is arclength-parametrized <strong>and</strong> κτ ≠0. Show that α has a Bertr<strong>and</strong> mate β if<strong>and</strong> only if there are constants r <strong>and</strong> c so that rκ + cτ =1.d. Given α, prove that if there is more than one curve β so that α <strong>and</strong> β are Bertr<strong>and</strong> mates,then there are <strong>in</strong>f<strong>in</strong>itely many such curves β <strong>and</strong> this occurs if <strong>and</strong> only if α is a circularhelix.20. (See Exercise 19.) Suppose α <strong>and</strong> β are Bertr<strong>and</strong> mates. Prove that the torsion of α <strong>and</strong> thetorsion of β at correspond<strong>in</strong>g po<strong>in</strong>ts have constant product.21. Suppose Y is a C 2 vector function on [a, b] with ‖Y‖ =1<strong>and</strong> Y, Y ′ , <strong>and</strong> Y ′′ everywhere l<strong>in</strong>early∫ t (<strong>in</strong>dependent. For any nonzero constant c, def<strong>in</strong>e α(t) =c Y(u)×Y ′ (u) ) du, t ∈ [a, b]. Provethat the curve α has constant torsion 1/c. (H<strong>in</strong>t: Show that B = ±Y.)22. a. Let α be an arclength-parametrized plane curve. We create a “parallel” curve β by tak<strong>in</strong>gβ = α + εN (for a fixed small positive value of ε). Expla<strong>in</strong> the term<strong>in</strong>ology <strong>and</strong> expressthe curvature of β <strong>in</strong> terms of ε <strong>and</strong> the curvature of α.b. Now let α be an arclength-parametrized space curve. Show that we can obta<strong>in</strong> a “parallel”curve β by tak<strong>in</strong>g β = α + ε ( (cos θ)N + (s<strong>in</strong> θ)B ) for an appropriate function θ. Howmany such parallel curves are there?c. Sketch such a parallel curve for a circular helix α.23. Suppose α is an arclength-parametrized curve, P = α(0), <strong>and</strong> κ(0) ≠0. Use Proposition 2.6to establish the follow<strong>in</strong>g:*a. Let Q = α(s) <strong>and</strong> R = α(t). Show that the plane spanned by P , Q, <strong>and</strong> R approachesthe osculat<strong>in</strong>g plane of α at P as s, t → 0.b. The osculat<strong>in</strong>g circle at P is the limit<strong>in</strong>g position of the circle pass<strong>in</strong>g through P , Q, <strong>and</strong>R as s, t → 0. Prove that the osculat<strong>in</strong>g circle has center Z = P + ( 1/κ(0) ) N(0) <strong>and</strong>radius 1/κ(0).c. The osculat<strong>in</strong>g sphere at P is the limit<strong>in</strong>g position of the sphere through P <strong>and</strong> threeneighbor<strong>in</strong>g po<strong>in</strong>ts on the curve, as the latter po<strong>in</strong>ts tend to P <strong>in</strong>dependently. Prove thata


22 Chapter 1. <strong>Curves</strong>the osculat<strong>in</strong>g sphere has center Z = P + ( 1/κ(0) ) N(0)+ ( 1/τ(0)(1/κ) ′ (0) ) B(0) <strong>and</strong> radius√(1/κ(0)) 2 +(1/τ(0)(1/κ) ′ (0)) 2 .d. How is the result of part c related to Exercise 18?24. a. Suppose β is a plane curve <strong>and</strong> C s is the circle centered at β(s) with radius r(s). Assum<strong>in</strong>gβ <strong>and</strong> r are differentiable functions, show that the circle C s is conta<strong>in</strong>ed <strong>in</strong>side the circleC t whenever t>sif <strong>and</strong> only if ‖β ′ (s)‖ ≤r ′ (s) for all s.b. Let α be arclength-parametrized plane curve <strong>and</strong> suppose κ is a decreas<strong>in</strong>g function. Provethat the osculat<strong>in</strong>g circle at α(s) lies <strong>in</strong>side the osculat<strong>in</strong>g circle at α(t) whenever t>s.(See Exercise 23 for the def<strong>in</strong>ition of the osculat<strong>in</strong>g circle.)25. Suppose the front wheel of a bicycle follows the arclength-parametrized plane curve α. Determ<strong>in</strong>ethe path β of the rear wheel, 1 unit away, as shown <strong>in</strong> Figure 2.8. (H<strong>in</strong>t: If the frontFigure 2.8wheel is turned an angle θ from the axle of the bike, start by writ<strong>in</strong>g α − β <strong>in</strong> terms of θ, T,<strong>and</strong> N. Your goal should be a differential equation that θ must satisfy. Note that the path ofthe rear wheel will obviously depend on the <strong>in</strong>itial condition θ(0). In all but the simplest ofcases, it may be impossible to solve the differential equation explicitly.)3. Some Global Results3.1. Space <strong>Curves</strong>. The fundamental notion <strong>in</strong> geometry (see Section 1 of the Appendix) isthat of congruence: When do two figures differ merely by a rigid motion? If the curve α ∗ is obta<strong>in</strong>edfrom the curve α by perform<strong>in</strong>g a rigid motion (composition of a translation <strong>and</strong> a rotation), thenthe Frenet frames at correspond<strong>in</strong>g po<strong>in</strong>ts differ by that same rigid motion, <strong>and</strong> the twist<strong>in</strong>g of theframes (which is what gives curvature <strong>and</strong> torsion) should be the same. (Note that a reflection willnot affect the curvature, but will change the sign of the torsion.)Theorem 3.1 (Fundamental Theorem of Curve Theory). Two space curves C <strong>and</strong> C ∗ are congruent(i.e., differ by a rigid motion) if <strong>and</strong> only if the correspond<strong>in</strong>g arclength parametrizationsα, α ∗ :[0,L] → R 3 have the property that κ(s) =κ ∗ (s) <strong>and</strong> τ(s) =τ ∗ (s) for all s ∈ [0,L].Proof. Suppose α ∗ =Ψ◦α for some rigid motion Ψ: R 3 → R 3 ,soΨ(x) =Ax + b for someb ∈ R 3 <strong>and</strong> some 3 × 3 orthogonal matrix A with det A > 0. Then α ∗ (s) = Aα(s) +b, so‖α ∗′ (s)‖ = ‖Aα ′ (s)‖ =1,s<strong>in</strong>ce A is orthogonal. Therefore, α ∗ is likewise arclength-parametrized,


§3. Some Global Results 23<strong>and</strong> T ∗ (s) = AT(s). Differentiat<strong>in</strong>g aga<strong>in</strong>, κ ∗ (s)N ∗ (s) = κ(s)AN(s). S<strong>in</strong>ce A is orthogonal,‖AN(s)‖ is a unit vector, <strong>and</strong> so N ∗ (s) =AN(s) <strong>and</strong> κ ∗ (s) =κ(s). But then B ∗ (s) =T ∗ (s) ×N ∗ (s) = AT(s) × AN(s) = A(T(s) × N(s)) = AB(s), <strong>in</strong>asmuch as orthogonal matrices maporthonormal bases to orthonormal bases <strong>and</strong> det A>0 <strong>in</strong>sures that orientation is preserved as well.Last, B ∗′ (s) =−τ ∗ (s)N ∗ (s) <strong>and</strong> B ∗′ (s) =AB ′ (s) =−τ(s)AN(s) =−τ(s)N ∗ (s), so τ ∗ (s) =τ(s),as required.Conversely, suppose κ = κ ∗ <strong>and</strong> τ = τ ∗ . Wenow def<strong>in</strong>e a rigid motion Ψ as follows. Letb = α ∗ (0) − α(0), <strong>and</strong> let A be the unique orthogonal matrix so that AT(0) = T ∗ (0), AN(0) =N ∗ (0), <strong>and</strong> AB(0) = B ∗ (0). A also has positive determ<strong>in</strong>ant, s<strong>in</strong>ce both orthonormal bases areright-h<strong>and</strong>ed. Set ˜α =Ψ◦α. Wenow claim that α ∗ (s) = ˜α(s) for all s ∈ [0,L]. Note, by ourargument <strong>in</strong> the first part of the proof, that ˜κ = κ = κ ∗ <strong>and</strong> ˜τ = τ = τ ∗ . Considerf(s) = ˜T(s) · T ∗ (s)+Ñ(s) · N∗ (s)+ ˜B(s) · B ∗ (s).We now differentiate f, us<strong>in</strong>g the Frenet formulas.f ′ (s) = ( ˜T′ (s) · T ∗ (s)+ ˜T(s) · T ∗′ (s) ) + ( Ñ ′ (s) · N ∗ (s)+Ñ(s) · N∗′ (s) )+ ( ˜B′ (s) · B ∗ (s)+ ˜B(s) · B ∗′ (s) )= κ(s) ( Ñ(s) · T ∗ (s)+ ˜T(s) · N ∗ (s) ) − κ(s) ( ˜T(s) · N ∗ (s)+Ñ(s) · T∗ (s) )=0,+ τ(s) ( ˜B(s) · N ∗ (s)+Ñ(s) · B∗ (s) ) − τ(s) ( Ñ(s) · B ∗ (s)+ ˜B(s) · N ∗ (s) )s<strong>in</strong>ce the first two terms cancel <strong>and</strong> the last two terms cancel. By construction, f(0) = 3, sof(s) =3forall s ∈ [0,L]. S<strong>in</strong>ce each of the <strong>in</strong>dividual dot products can be at most 1, the onlyway the sum can be 3 for all s is for each to be 1 for all s, <strong>and</strong> this <strong>in</strong> turn can happen onlywhen ˜T(s) =T ∗ (s), Ñ(s) =N ∗ (s), <strong>and</strong> ˜B(s) =B ∗ (s) for all s ∈ [0,L]. In particular, s<strong>in</strong>ce˜α ′ (s) = ˜T(s) =T ∗ (s) =α ∗′ (s) <strong>and</strong> ˜α(0) = α ∗ (0), it follows that ˜α(s) =α ∗ (s) for all s ∈ [0,L], aswe wished to show. □Remark. The latter half of this proof can be replaced by assert<strong>in</strong>g the uniqueness of solutionsof a system of differential equations, as we will see <strong>in</strong> a moment. Also see Exercise A.3.1 for amatrix-computational version of the proof we just did.Example 1. We now see that the only curves with constant κ <strong>and</strong> τ are circular helices.▽Perhaps more <strong>in</strong>terest<strong>in</strong>g is the existence question: Given cont<strong>in</strong>uous functions κ, τ :[0,L] → R(with κ everywhere positive), is there a space curve with those as its curvature <strong>and</strong> torsion? Theanswer is yes, <strong>and</strong> this is an immediate consequence of the fundamental existence theorem fordifferential equations, Theorem 3.1 of the Appendix. That is, we let⎡⎤⎡⎤| | |0 −κ(s) 0⎢F (s) = ⎣T(s) N(s) B(s)| | |⎥⎢⎦ <strong>and</strong> K(s) = ⎣κ(s) 0 −τ(s)0 τ(s) 0Then <strong>in</strong>tegrat<strong>in</strong>g the l<strong>in</strong>ear system of ord<strong>in</strong>ary differential equations F ′ (s) =F (s)K(s), F (0) = F 0 ,gives us the Frenet frame everywhere along the curve, <strong>and</strong> we recover α by <strong>in</strong>tegrat<strong>in</strong>g T(s).⎥⎦ .


24 Chapter 1. <strong>Curves</strong>We turn now to the concept of total curvature of a closed space curve, which is the <strong>in</strong>tegral of itscurvature. That is, if α: [0,L] → R 3 is an arclength-parametrized curve with α(0) = α(L), its totalcurvature is∫ L0κ(s)ds. This quantity can be <strong>in</strong>terpreted geometrically as follows: The Gauss mapof α is the map to the unit sphere, Σ, given by the unit tangent vector T: [0,L] → Σ; its image,Γ, is classically called the tangent <strong>in</strong>dicatrix of α. Observe that—provided the Gauss map is one-TFigure 3.1to-one—the length of Γ is the total curvature of α, s<strong>in</strong>ce length(Γ) =∫ L0‖T ′ (s)‖ds =∫ L0κ(s)ds.A prelim<strong>in</strong>ary question to ask is this: What curves Γ <strong>in</strong> the unit sphere can be the Gauss map ofsome closed space curve α? S<strong>in</strong>ce α(s) =α(0) +condition is that∫ L0∫ s0T(u)du, wesee that a necessary <strong>and</strong> sufficientT(s)ds = 0. (Note, however, that this depends on the arclength parametrizationof the orig<strong>in</strong>al curve <strong>and</strong> is not a parametrization-<strong>in</strong>dependent condition on the image curveΓ ⊂ Σ.) We do, nevertheless, have the follow<strong>in</strong>g geometric consequence of this condition. For any(unit) vector A, wehave0=A ·∫ L0T(s)ds =∫ L0(T(s) · A)ds,<strong>and</strong> so the average value of T · A must be 0. In particular, the tangent <strong>in</strong>dicatrix must cross thegreat circle with normal vector A. That is, if the curve Γ is to be a tangent <strong>in</strong>dicatrix, it must be“balanced” with respect to every direction A. Itisnatural to ask for the shortest curve(s) withthis property.If ξ ∈ Σ, let ξ ⊥ denote the (oriented) great circle with normal vector ξ.Proposition 3.2 (Crofton’s formula). Let Γ be apiecewise-C 1 curve on the sphere. Thenlength(Γ) = 1 ∫#(Γ ∩ ξ ⊥ )dξ4 Σ= π × (the average number of <strong>in</strong>tersections of Γ with all great circles).(Here dξ represents the usual element of surface area on Σ.)Proof. We leave this to the reader <strong>in</strong> Exercise 11.□Remark. Although we don’t stop to justify it here, the set of ξ for which #(Γ ∩ ξ ⊥ )is<strong>in</strong>f<strong>in</strong>iteis a set of measure zero, <strong>and</strong> so the <strong>in</strong>tegral makes sense.


§3. Some Global Results 25Apply<strong>in</strong>g this to the case of the tangent <strong>in</strong>dicatrix of a closed space curve, we deduce thefollow<strong>in</strong>g classical result.Theorem 3.3 (Fenchel). The total curvature of any closed space curve is at least 2π, <strong>and</strong>equality holds if <strong>and</strong> only if the curve is a convex, planar curve.Proof. Let Γ be the tangent <strong>in</strong>dicatrix of our space curve. ∫ For almost every ξ ∈ Σwehave#(Γ ∩ ξ ⊥ ) ≥ 2, <strong>and</strong> so it follows from Proposition 3.2 that κds = length(Γ) ≥ 1 (2)(4π) =2π.4Now, we claim that if Γ is a connected, closed curve <strong>in</strong> Σ of length ≤ 2π, then Γ lies <strong>in</strong> a closedhemisphere. It will follow, then, that if Γ is a tangent <strong>in</strong>dicatrix of length 2π, itmust be a greatcircle. (For if Γ lies <strong>in</strong> the hemisphere A · x ≥ 0,∫ Lgreat circle A · x = 0.) It follows that the curve is planar.0T(s) · Ads =0forces T · A =0,soΓis thePNQΓ 1Γ 1Figure 3.2To prove the claim, we proceed as follows. Suppose length(Γ) ≤ 2π. Choose P <strong>and</strong> Q <strong>in</strong> Γso that the arcs Γ 1 = PQ ⌢<strong>and</strong> Γ 2 = QP ⌢have the same length. Choose N bisect<strong>in</strong>g the shortergreat circle arc from P to Q, asshown <strong>in</strong> Figure 3.2. For convenience, we rotate the picture sothat N is the north pole of the sphere. Suppose now that the curve Γ 1 were to enter the southernhemisphere; let Γ 1 denote the reflection of Γ 1 across the north pole (follow<strong>in</strong>g arcs of great circlethrough N). Now, Γ 1 ∪ Γ 1 is a closed curve conta<strong>in</strong><strong>in</strong>g a pair of antipodal po<strong>in</strong>ts <strong>and</strong> therefore islonger than a great circle. (See Exercise 1.) S<strong>in</strong>ce Γ 1 ∪ Γ 1 has the same length as Γ, we see thatlength(Γ) > 2π, which is a contradiction. Therefore Γ <strong>in</strong>deed lies <strong>in</strong> the northern hemisphere. □We now sketch the proof of a result that has led to many <strong>in</strong>terest<strong>in</strong>g questions <strong>in</strong> higherdimensions. We say a simple closed space curve is knotted if we cannot fill it <strong>in</strong> with a disk.Theorem 3.4 (Fary-Milnor). If a simple closed space curve is knotted, then its total curvatureis at least 4π.Sketch of proof. Suppose the total curvature of C is less than 4π. Then the average number#(Γ∩ξ ⊥ ) < 4. S<strong>in</strong>ce this is generically an even number ≥ 2 (whenever the great circle isn’t tangentto Γ), there must be an open set of ξ’s for which we have #(Γ∩ξ ⊥ )=2. Choose one such, ξ 0 . Thismeans that the tangent vector to C is only perpendicular to ξ 0 twice, so the function f(x) =x · ξ 0has only two critical po<strong>in</strong>ts. That is, the planes perpendicular to ξ 0 will (by Rolle’s Theorem)


26 Chapter 1. <strong>Curves</strong><strong>in</strong>tersect C either <strong>in</strong> a l<strong>in</strong>e segment or <strong>in</strong> a s<strong>in</strong>gle po<strong>in</strong>t (at the top <strong>and</strong> bottom); that is, by mov<strong>in</strong>gthese planes from the bottom of C to the top, we fill <strong>in</strong> a disk, so C is unknotted. □3.2. Plane <strong>Curves</strong>. We conclude this chapter with some results on plane curves. <strong>First</strong>, itmakes sense to consider curvature with a sign. Rather than proceed<strong>in</strong>g as we did <strong>in</strong> Section 2,<strong>in</strong>stead, given an arclength-parametrized curve α, def<strong>in</strong>e N(s) sothat {T(s), N(s)} is a righth<strong>and</strong>edbasis for R 2 (i.e., one turns counterclockwise from T(s) toN(s)), <strong>and</strong> then set T ′ (s) =κ(s)N(s), as before. So κ>0 when T is twist<strong>in</strong>g counterclockwise <strong>and</strong> κ0}, U 2 = {(x, y) ∈ S 1 :x0}, <strong>and</strong> U 4 = {(x, y) ∈ S 1 : y


§3. Some Global Results 27Def<strong>in</strong>e θ(0) so that T(0) = ( cos θ(0), s<strong>in</strong> θ(0) ) . Let S = {s ∈ [0,L]:θ is def<strong>in</strong>ed <strong>and</strong> C 1 on[0,s]}, <strong>and</strong> let s 0 = sup S. Suppose first that s 0 0sothat0T(s) ∈ U i for all s with |s − s 0 |


28 Chapter 1. <strong>Curves</strong>h(s,t)α(t)α(s)α(0)T(0)Figure 3.4Recall that one of the ways of characteriz<strong>in</strong>g a convex function f : R → R is that its graph lieon one side of each of its tangent l<strong>in</strong>es. So we make the follow<strong>in</strong>gDef<strong>in</strong>ition. The regular closed plane curve α is convex if it lies on one side of its tangent l<strong>in</strong>eat each po<strong>in</strong>t.Proposition 3.8. A simple closed regular plane curve C is convex if <strong>and</strong> only if we can choosethe orientation of the curve so that κ ≥ 0 everywhere.Remark. We leave it to the reader <strong>in</strong> Exercise 2 to give a non-simple closed curve for whichthis result is false.Proof. Assume, without loss of generality, that T(0) =(1, 0) <strong>and</strong> the curve is oriented counterclockwise.Us<strong>in</strong>g the function θ constructed <strong>in</strong> Lemma 3.6, the condition that κ ≥ 0isequivalentto the condition that θ is a nondecreas<strong>in</strong>g function with θ(L) =2π.Suppose first that θ is nondecreas<strong>in</strong>g <strong>and</strong> C is not convex. Then we can f<strong>in</strong>d a po<strong>in</strong>t P = α(s 0 )on the curve <strong>and</strong> values s ′ 1 , s′ 2 so that α(s′ 1 ) <strong>and</strong> α(s′ 2 ) lie on opposite sides of the tangent l<strong>in</strong>e to Cat P . Then, by the maximum value theorem, there are values s 1 <strong>and</strong> s 2 so that α(s 1 )isthe greatestdistance “above” the tangent l<strong>in</strong>e <strong>and</strong> α(s 2 )isthe greatest distance “below.” Consider the unittangent vectors T(s 0 ), T(s 1 ), <strong>and</strong> T(s 2 ). S<strong>in</strong>ce these vectors are either parallel or anti-parallel,some pair must be identical. Lett<strong>in</strong>g the respective values of s be s ∗ <strong>and</strong> s ∗∗ with s ∗


§3. Some Global Results 29at M. Thus, it must be that PQ ⊂ C, soθ(s) =θ(s 1 )=θ(s 2 ) for all s ∈ [s 1 ,s 2 ]. Therefore, θ isnondecreas<strong>in</strong>g, <strong>and</strong> we are done. □Def<strong>in</strong>ition. A critical po<strong>in</strong>t of κ is called a vertex of the curve C.A closed curve must have at least two vertices: the maximum <strong>and</strong> m<strong>in</strong>imum po<strong>in</strong>ts of κ. Everypo<strong>in</strong>t of a circle is a vertex. We conclude with the follow<strong>in</strong>gProposition 3.9 (Four Vertex Theorem). A closed convex plane curve has at least four vertices.Proof. Suppose that C has fewer than four vertices. As we see from Figure 3.5, either κ musthave two critical po<strong>in</strong>ts (maximum <strong>and</strong> m<strong>in</strong>imum) or κ must have three critical po<strong>in</strong>ts (maximum,0 L 0 LFigure 3.5m<strong>in</strong>imum, <strong>and</strong> <strong>in</strong>flection po<strong>in</strong>t). More precisely, suppose that κ <strong>in</strong>creases from P to Q <strong>and</strong> decreasesfrom Q to P . Without loss of generality, let P be the orig<strong>in</strong> <strong>and</strong> suppose ∫ the equation of PQ ←→ isA · x =0. Choose A so that κ ′ (s) ≥ 0 precisely when A · α(s) ≥ 0. Then κ ′ (s)(A · α(s))ds > 0.Let à be the vector obta<strong>in</strong>ed by rotat<strong>in</strong>g A through an angle of π/2. Then, <strong>in</strong>tegrat<strong>in</strong>g by parts,we have∫∫∫− κ ′ (s)(A · α(s))ds = κ(s)(A · T(s))ds = κ(s)(à · N(s))dsCCC∫∫= à · κ(s)N(s)ds = à · T ′ (s)ds =0.From this contradiction, we <strong>in</strong>fer that C must have at least four vertices.C3.3. The Isoperimetric Inequality. One of the classic questions <strong>in</strong> mathematics is thefollow<strong>in</strong>g: Given a closed curve of length L, what shape will enclose the most area? A littleexperimentation will most likely lead the reader to theTheorem 3.10 (Isoperimetric Inequality). If a simple closed plane curve C has length L <strong>and</strong>encloses area A, then<strong>and</strong> equality holds if <strong>and</strong> only if C is a circle.L 2 ≥ 4πA,CC□


30 Chapter 1. <strong>Curves</strong>y( x(s) ,y(s))α(0)Cα(s 0 )( x(s),y(s) )xl 1 l 2CRFigure 3.6Proof. There are a number of different proofs, but we give one (due to E. Schmidt, 1939) basedon Green’s Theorem, Theorem 2.6 of the Appendix, <strong>and</strong>—not surpris<strong>in</strong>gly—rely<strong>in</strong>g heavily on thegeometric-arithmetic mean <strong>in</strong>equality <strong>and</strong> the Cauchy-Schwarz <strong>in</strong>equality (see Exercise A.1.2). Wechoose parallel l<strong>in</strong>es l 1 <strong>and</strong> l 2 tangent to, <strong>and</strong> enclos<strong>in</strong>g, C, aspictured <strong>in</strong> Figure 3.6. We drawa circle C of radius R with those same tangent l<strong>in</strong>es <strong>and</strong> put the orig<strong>in</strong> at its center, with they-axis parallel to l i .Wenow parametrize C by arclength by α(s) =(x(s),y(s)), s ∈ [0,L], tak<strong>in</strong>gα(0) ∈ l 1 <strong>and</strong> α(s 0 ) ∈ l 2 .Wethen consider α: [0,L] → R 2 given by⎧α(s) = ( x(s), y(s) ) ⎨( √x(s), − R 2 − x(s) 2) , 0 ≤ s ≤ s 0= ( √⎩ x(s), R 2 − x(s) 2) , s 0 ≤ s ≤ L .(α needn’t be a parametrization of the circle C, s<strong>in</strong>ce it may cover certa<strong>in</strong> portions multiple times,but that’s no problem.) Lett<strong>in</strong>g A denote the area enclosed by C <strong>and</strong> A = πR 2 that enclosed byC, wehave (by Exercise A.2.5)A =∫ LA = πR 2 = −0∫ L0x(s)y ′ (s)dsy(s)x ′ (s)ds = −∫ L0y(s)x ′ (s)ds.Add<strong>in</strong>g these equations <strong>and</strong> apply<strong>in</strong>g the Cauchy-Schwarz <strong>in</strong>equality, we have∫ LA + πR 2 (= x(s)y ′ (s) − y(s)x ′ (s) ) ∫ L ( ) (ds = x(s), y(s) · y ′ (s), −x ′ (s) ) ds(∗)≤0∫ L0‖ ( x(s), y(s) ) ‖‖ ( y ′ (s), −x ′ (s) ) ‖ds = RL,<strong>in</strong>asmuch as ‖(y ′ (s), −x ′ (s))‖ = ‖(x ′ (s),y ′ (s))‖ =1s<strong>in</strong>ce α is arclength-parametrized. We nowrecall the arithmetic-geometric mean <strong>in</strong>equality:√ a + b ab ≤ for positive numbers a <strong>and</strong> b,20


§3. Some Global Results 31with equality hold<strong>in</strong>g if <strong>and</strong> only if a = b. Wetherefore have√A√πR 2 ≤ A + πR22≤ RL2 ,so 4πA ≤ L 2 .Now suppose equality holds here. Then we must have A = πR 2 <strong>and</strong> L =2πR. Itfollows thatthe curve C has the same breadth <strong>in</strong> all directions (s<strong>in</strong>ce L now determ<strong>in</strong>es R). But equality mustalso hold <strong>in</strong> (∗), so the vectors α(s) = ( x(s), y(s) ) <strong>and</strong> ( y ′ (s), −x ′ (s) ) must be everywhere parallel.S<strong>in</strong>ce the first vector has length R <strong>and</strong> the second has length 1, we <strong>in</strong>fer that(x(s), y(s))= R(y ′ (s), −x ′ (s) ) ,<strong>and</strong> so x(s) =Ry ′ (s). By our remark at the beg<strong>in</strong>n<strong>in</strong>g of this paragraph, the same result will holdif we rotate the axes π/2; let y = y 0 be the l<strong>in</strong>e halfway between the enclos<strong>in</strong>g horizontal l<strong>in</strong>es l i .Now, substitut<strong>in</strong>g y − y 0 for x <strong>and</strong> −x for y, sowehavey(s) − y 0 = −Rx ′ (s), as well. Therefore,x(s) 2 + ( ) 2y(s) − y 0 = R 2 (x ′ (s) 2 + y ′ (s) 2 )=R 2 , <strong>and</strong> C is <strong>in</strong>deed a circle of radius R. □EXERCISES 1.31. a. Prove that the shortest path between two po<strong>in</strong>ts on the unit sphere is the arc of a greatcircle connect<strong>in</strong>g them. (H<strong>in</strong>t: Without loss of generality, take one po<strong>in</strong>t to be (0, 0, 1)<strong>and</strong> the other to be (s<strong>in</strong> u 0 , 0, cos u 0 ). Let α(t) =(s<strong>in</strong> u(t) cos v(t), s<strong>in</strong> u(t) s<strong>in</strong> v(t), cos u(t)),a ≤ t ≤ b, beanarbitrary path with u(a) =0,v(a) =0,u(b) =u 0 , v(b) =0, calculate thearclength of α, <strong>and</strong> show that it is smallest when v(t) =0for all t.)b. Prove that if P <strong>and</strong> Q are po<strong>in</strong>ts on the unit sphere, then the shortest path between themhas length arccos(P · Q).2. Give a closed plane curve C with κ>0 that is not convex.3. Draw closed plane curves with rotation <strong>in</strong>dices 0, 2, −2, <strong>and</strong> 3, respectively.*4. Suppose C is a simple closed plane curve with 0


32 Chapter 1. <strong>Curves</strong>ɛ2ɛ 1ɛ 3CFigure 3.78. Let C be a C 2 closed space curve, say parametrized by arclength by α: [0,L] → R 3 . A unitnormal field X on C is a C 1 vector-valued function with X(0) = X(L) <strong>and</strong> X(s) · T(s) =0<strong>and</strong>‖X(s)‖ =1for all s. Wedef<strong>in</strong>e the twist of X to betw(C, X) = 12π∫ L0X ′ (s) · (T(s) × X(s))ds.a. Show that if X <strong>and</strong> X ∗ are two unit normal fields on C, then tw(C, X) <strong>and</strong> tw(C, X ∗ )differ by an <strong>in</strong>teger. The fractional part of tw(C, X) (i.e., the twist mod 1) is called thetotal twist of C. (H<strong>in</strong>t: Write X(s) =cos θ(s)N(s)+s<strong>in</strong> θ(s)B(s).) ∫1b. Prove that the total twist of C equals the fractional part of τds.2π Cc. Prove that if a closed curve lies on a sphere, then its total twist is 0. (H<strong>in</strong>t: Choose anobvious c<strong>and</strong>idate for X.)Remark. W. Scherrer proved <strong>in</strong> 1940 that if the total twist of every closed curve on asurface is 0, then that surface must be a (subset of a) plane or sphere.9. A convex plane curve can be determ<strong>in</strong>ed by its tangent l<strong>in</strong>es (cos θ)x+(s<strong>in</strong> θ)y = p(θ), called itssupport l<strong>in</strong>es. The function p(θ) iscalled the support function. (Here θ is the polar coord<strong>in</strong>ate.)a. Prove that the l<strong>in</strong>e given above is tangent to the curve at the po<strong>in</strong>tα(θ) =(p(θ) cos θ − p ′ (θ) s<strong>in</strong> θ, p(θ) s<strong>in</strong> θ + p ′ (θ) cos θ).b. Prove that the curvature of the curve at α(θ) is1 /( p(θ)+p ′′ (θ) ) .c. Prove that the length of α is given by L =∫ 2π0p(θ)dθ.∫ 2πd. Prove that the area enclosed by α is given by A = 1 (p(θ) 2 − p ′ (θ) 2) dθ.2 0e. Use the answer to part c to reprove the result of Exercise 7.10. (See Exercise 1.2.22.) Under what circumstances does a closed space curve have a parallel curvethat is also closed? (H<strong>in</strong>t: Exercise 8 should be relevant.)11. Prove Proposition 3.2 as follows. Let α: [0,L] → Σbethe arclength parametrization of Γ,<strong>and</strong> def<strong>in</strong>e F: [0,L] × [0, 2π) → ΣbyF(s, φ) =ξ, where ξ ⊥ is the great circle mak<strong>in</strong>g angle φwith Γ at α(s). Check that F takes on the value ξ precisely #(Γ ∩ ξ ⊥ ) times, so that F is a


§3. Some Global Results 33ξΓ⊥ξvα(s)TφvFigure 3.8“multi-parametrization” of Σ that gives us∫∫ L ∫ 2π#(Γ ∩ ξ ⊥ )dξ =∂F∥Σ0 0 ∂s × ∂F∂φ∥ dφds.Compute that∂F∥ ∂s × ∂F∂φ∥ = | s<strong>in</strong> φ| (this is the hard part) <strong>and</strong> f<strong>in</strong>ish the proof. (H<strong>in</strong>ts: Aspictured <strong>in</strong> Figure 3.8, show v(s, φ) =cos φT(s)+s<strong>in</strong> φ(α(s) × T(s)) is the tangent vector tothe great circle ξ ⊥ <strong>and</strong> deduce that F(s, φ) =α(s) × v(s, φ). Show that ∂F ∂v<strong>and</strong> α ×∂φ ∂s areboth multiples of v.)12. Complete the details of the proof of the <strong>in</strong>dicated step <strong>in</strong> the proof of Theorem 3.5, as follows.Pick an <strong>in</strong>terior po<strong>in</strong>t s 0 ∈ ∆.a. Choose ˜θ(s 0 )sothat h(s 0 )= ( cos ˜θ(s 0 ), s<strong>in</strong> ˜θ(s 0 ) ) . Use the procedure of the proof ofLemma 3.6 to determ<strong>in</strong>e ˜θ uniquely as a function that is cont<strong>in</strong>uous on each ray −→ s 0 s forevery s ∈ ∆.b. Choose s 1 ∈ ∆. Show first (us<strong>in</strong>g the fact that a cont<strong>in</strong>uous function on the <strong>in</strong>terval s 0 s 1is uniformly cont<strong>in</strong>uous) that there is δ>0sothat whenever ‖s − s ′ ‖ 0, show that there is a neighborhood U of s 1 so that whenever s ∈ U we have‖s − s 1 ‖


CHAPTER 2<strong>Surfaces</strong>: Local Theory1. Parametrized <strong>Surfaces</strong> <strong>and</strong> the <strong>First</strong> Fundamental FormLet U be an open set <strong>in</strong> R 2 .Afunction f : U → R m (for us, m =1<strong>and</strong> 3 will be most common)is called C 1 if f <strong>and</strong> its partial derivatives ∂f ∂f<strong>and</strong> are all cont<strong>in</strong>uous. We will ord<strong>in</strong>arily use∂u ∂v(u, v) ascoord<strong>in</strong>ates <strong>in</strong> our parameter space, <strong>and</strong> (x, y, z) ascoord<strong>in</strong>ates <strong>in</strong> R 3 . Similarly, for anyk ≥ 2, we say f is C k if all its partial derivatives of order up to k exist <strong>and</strong> are cont<strong>in</strong>uous. We say fis smooth if f is C k for every positive <strong>in</strong>teger k. Wewill henceforth assume all our functions are C kfor k ≥ 3. One of the crucial results for differential geometry is that if f is C 2 , then(<strong>and</strong> similarly for higher-order derivatives).Notation: We will often also use subscripts to <strong>in</strong>dicate partial derivatives, as follows:f uf vf uuf uv =(f u ) v↔↔↔↔∂f∂u∂f∂v∂ 2 f∂u 2∂ 2 f∂v∂u∂2 f∂u∂v =Def<strong>in</strong>ition. A regular parametrization of a subset M ⊂ R 3 isa(C 3 ) one-to-one functionx: U → M ⊂ R 3 so that x u × x v ≠ 0∂2 f∂v∂ufor some open set U ⊂ R 2 . 1 A connected subset M ⊂ R 3 is called a surface if each po<strong>in</strong>t has aneighborhood that is regularly parametrized.We might consider the curves on M obta<strong>in</strong>ed by fix<strong>in</strong>g v = v 0 <strong>and</strong> vary<strong>in</strong>g u, called a u-curve,<strong>and</strong> obta<strong>in</strong>ed by fix<strong>in</strong>g u = u 0 <strong>and</strong> vary<strong>in</strong>g v, called a v-curve; these are depicted <strong>in</strong> Figure 1.1. Atthe po<strong>in</strong>t P = x(u 0 ,v 0 ), we see that x u (u 0 ,v 0 )istangent to the u-curve <strong>and</strong> x v (u 0 ,v 0 )istangentto the v-curve. We are requir<strong>in</strong>g that these vectors span a plane, whose normal vector is given byx u × x v .Example 1. We give some basic examples of parametrized surfaces. Note that our parametersdo not necessarily range over an open set of values.1 For technical reasons with which we shall not concern ourselves <strong>in</strong> this course, we should also require that the<strong>in</strong>verse function x −1 : x(U) → U be cont<strong>in</strong>uous.34


§1. Parametrized <strong>Surfaces</strong> <strong>and</strong> the <strong>First</strong> Fundamental Form 35vxx vx u ×x vv-curvex uuu-curveFigure 1.1(a) The graph of a function f : U → R, z = f(x, y), is parametrized by x(u, v) =(u, v, f(u, v)).Note that x u × x v =(−f u , −f v , 1) ≠ 0, sothis is always a regular parametrization.(b) The helicoid, asshown <strong>in</strong> Figure 1.2, is the surface formed by draw<strong>in</strong>g horizontal rays fromFigure 1.2the axis of the helix α(t) =(cos t, s<strong>in</strong> t, bt) topo<strong>in</strong>ts on the helix:x(u, v) =(u cos v, u s<strong>in</strong> v, bv), u ≥ 0, v∈ R.Note that x u × x v =(b s<strong>in</strong> v, −b cos v, u) ≠ 0. The u-curves are rays <strong>and</strong> the v-curves arehelices.(c) The torus (surface of a doughnut) is formed by rotat<strong>in</strong>g a circle of radius b about a circleof radius a > b ly<strong>in</strong>g <strong>in</strong> an orthogonal plane, as pictured <strong>in</strong> Figure 1.3. The regularparametrization is given byx(u, v) =((a + b cos u) cos v, (a + b cos u) s<strong>in</strong> v, b s<strong>in</strong> u), 0 ≤ u, v < 2π.Then x u × x v = −b(a + b cos u) ( cos u cos v, cos u s<strong>in</strong> v, s<strong>in</strong> u ) , which is never 0.(d) The st<strong>and</strong>ard parametrization of the unit sphere Σ is given by spherical coord<strong>in</strong>ates (φ, θ) ↔(u, v):x(u, v) =(s<strong>in</strong> u cos v, s<strong>in</strong> u s<strong>in</strong> v, cos u),0


36 Chapter 2. <strong>Surfaces</strong>: Local TheoryavubuFigure 1.3S<strong>in</strong>ce x u × x v = s<strong>in</strong> u(s<strong>in</strong> u cos v, s<strong>in</strong> u s<strong>in</strong> v, cos u) =(s<strong>in</strong> u)x(u, v), the parametrization isregular away from u =0,π, which we’ve excluded anyhow because x fails to be one-to-oneat such po<strong>in</strong>ts. The u-curves are the so-called l<strong>in</strong>es of longitude <strong>and</strong> the v-curves are thel<strong>in</strong>es of latitude on the sphere.(e) Another <strong>in</strong>terest<strong>in</strong>g parametrization of the sphere is given by stereographic projection. (Cf.Exercise 1.1.1.) We parametrize the unit sphere less the north pole (0, 0, 1) by the xy-plane,(0,0,1)(x,y,z)(u,v,0)Figure 1.4assign<strong>in</strong>g to each (u, v) the po<strong>in</strong>t (≠ (0, 0, 1)) where the l<strong>in</strong>e through (0, 0, 1) <strong>and</strong> (u, v, 0)<strong>in</strong>tersects the unit sphere, as pictured <strong>in</strong> Figure 1.4. We leave it to the reader to derive thefollow<strong>in</strong>g formula <strong>in</strong> Exercise 1:(x(u, v) =2uu 2 + v 2 +1 ,2vu 2 + v 2 +1 , u2 + v 2 )− 1u 2 + v 2 . ▽+1For our last examples, we give two general classes of surfaces that will appear throughout ourwork.


§1. Parametrized <strong>Surfaces</strong> <strong>and</strong> the <strong>First</strong> Fundamental Form 37Example 2. Let I ⊂ R be an <strong>in</strong>terval, <strong>and</strong> let α(u) =(0,f(u),g(u)), u ∈ I, bearegularparametrized plane curve with f>0. Then the surface of revolution obta<strong>in</strong>ed by rotat<strong>in</strong>g α aboutthe z-axis is parametrized byx(u, v) = ( f(u) cos v, f(u) s<strong>in</strong> v, g(u) ) ,u ∈ I, 0 ≤ v


38 Chapter 2. <strong>Surfaces</strong>: Local Theoryhypothesis, the vectors x u <strong>and</strong> x v are l<strong>in</strong>early <strong>in</strong>dependent <strong>and</strong> must therefore span a plane. Wenow make this an officialDef<strong>in</strong>ition. Let M be a regular parametrized surface, <strong>and</strong> let P ∈ M. Then choose a regularparametrization x: U → M ⊂ R 3 with P = x(u 0 ,v 0 ). We def<strong>in</strong>e the tangent plane of M at P tobe the subspace T P M spanned by x u <strong>and</strong> x v (evaluated at (u 0 ,v 0 )).Remark. The alert reader may wonder what happens if two people pick two different such localparametrizations of M near P .Dothey both provide the same plane T P M? This sort of questionis very common <strong>in</strong> differential geometry, <strong>and</strong> is not one we <strong>in</strong>tend to belabor <strong>in</strong> this <strong>in</strong>troductorycourse. However, to get a feel for how such arguments go, the reader may work Exercise 11.There are two unit vectors orthogonal to the tangent plane T P M. Given a regular parametrizationx, weknow that x u × x v is a nonzero vector orthogonal to the plane spanned by x u <strong>and</strong> x v ;we obta<strong>in</strong> the correspond<strong>in</strong>g unit vector by tak<strong>in</strong>gn =x u × x v‖x u × x v ‖ .This is called the unit normal of the parametrized surface.Example 4. We know from basic geometry <strong>and</strong> vector calculus that the unit normal of theunit sphere centered at the orig<strong>in</strong> should be the position vector itself. This is <strong>in</strong> fact what wediscovered <strong>in</strong> Example 1(d). ▽Example 5. Consider the helicoid given <strong>in</strong> Example 1(b). Then, as we saw, x u × x v =1(b s<strong>in</strong> v, −b cos v, u), <strong>and</strong> n = √u 2 + b (b s<strong>in</strong> v, −b cos v, u). As we move along a rul<strong>in</strong>g v = v 0,2the normal starts horizontal at u =0(where the surface becomes vertical) <strong>and</strong> rotates <strong>in</strong> the planeorthogonal to the rul<strong>in</strong>g, becom<strong>in</strong>g more <strong>and</strong> more vertical as we move out the rul<strong>in</strong>g. ▽We saw <strong>in</strong> Chapter 1 that the geometry of a space curve is best understood by calculat<strong>in</strong>g (atleast <strong>in</strong> pr<strong>in</strong>ciple) with an arclength parametrization. It would be nice, analogously, if we couldf<strong>in</strong>d a parametrization x(u, v) ofasurface so that x u <strong>and</strong> x v form an orthonormal basis at eachpo<strong>in</strong>t. We’ll see later that this can happen only very rarely. But it makes it natural to <strong>in</strong>troducewhat is classically called the first fundamental form, I P (U, V) =U · V, for U, V ∈ T P M.Work<strong>in</strong>g<strong>in</strong> a parametrization, we have the natural basis {x u , x v }, <strong>and</strong> so we def<strong>in</strong>eE =I P (x u , x u )=x u · x uF =I P (x u , x v )=x u · x v = x v · x u =I P (x v , x u )G =I P (x v , x v )=x v · x v ,<strong>and</strong> it is often convenient to put these <strong>in</strong> as entries of a (symmetric) matrix:[ ]E FI P = .F GThen, given tangent vectors U = ax u + bx v <strong>and</strong> V = cx u + dx v ∈ T P M,wehaveU · V =I P (U, V) =(ax u + bx v ) · (cx u + dx v )=E(ac)+F (ad + bc)+G(bd).


§1. Parametrized <strong>Surfaces</strong> <strong>and</strong> the <strong>First</strong> Fundamental Form 39In particular, ‖U‖ 2 =I P (U, U) =Ea 2 +2Fab+ Gb 2 .Suppose M <strong>and</strong> M ∗ are surfaces. We say they are locally isometric if for each P ∈ M there are aregular parametrization x: U → M with x(u 0 ,v 0 )=P <strong>and</strong> a regular parametrization x ∗ : U → M ∗(us<strong>in</strong>g the same doma<strong>in</strong> U ⊂ R 2 ) with the property that I P =I ∗ P ∗ whenever P = x(u, v) <strong>and</strong>P ∗ = x ∗ (u, v) for some (u, v) ∈ U. That is, the function f = x ∗ ◦x −1 : x(U) → x ∗ (U) isaone-to-onecorrespondence that preserves the first fundamental form <strong>and</strong> is therefore distance-preserv<strong>in</strong>g (seeExercise 3).Example 6. Parametrize a portion of the plane (say, a piece of paper) by x(u, v) =(u, v, 0) <strong>and</strong>aportion of a cyl<strong>in</strong>der by x ∗ (u, v) =(cos u, s<strong>in</strong> u, v). Then it is easy to calculate that E = E ∗ =1,F = F ∗ =0,<strong>and</strong> G = G ∗ =1,sothese surfaces, pictured <strong>in</strong> Figure 1.6, are locally isometric. OnfoldsealFigure 1.6the other h<strong>and</strong>, if we let u vary from 0 to 2π, the rectangle <strong>and</strong> the cyl<strong>in</strong>der are not globally isometricbecause po<strong>in</strong>ts far away <strong>in</strong> the rectangle can become very close (or identical) <strong>in</strong> the cyl<strong>in</strong>der. ▽If α(t) =x(u(t),v(t)) is a curve on the parametrized surface M with α(t 0 )=x(u 0 ,v 0 )=P ,then it is an immediate consequence of the cha<strong>in</strong> rule, Theorem 2.2 of the Appendix, thatα ′ (t 0 )=u ′ (t 0 )x u (u 0 ,v 0 )+v ′ (t 0 )x v (u 0 ,v 0 ).(Customarily we will write simply x u , the po<strong>in</strong>t (u 0 ,v 0 )atwhich it is evaluated be<strong>in</strong>g assumed.)That is, if the tangent vector (u ′ (t 0 ),v ′ (t 0 )) back <strong>in</strong> the “parameter space” is (a, b), then the tangentvector to α at P is the correspond<strong>in</strong>g l<strong>in</strong>ear comb<strong>in</strong>ation ax u + bx v .Infancy terms, this is merelya consequence of the l<strong>in</strong>earity of the derivative of x. Wesayaparametrization x(u, v) isconformal(a,b)xPx v(u 0 ,v 0 )x uax u + bx vFigure 1.7


40 Chapter 2. <strong>Surfaces</strong>: Local Theoryif angles measured <strong>in</strong> the uv-plane agree with correspond<strong>in</strong>g angles <strong>in</strong> T P M for all P .Weleave itto the reader to check <strong>in</strong> Exercise 5 that this is equivalent to the conditions E = G, F =0.S<strong>in</strong>cewe have[EF]F=G[]x u · x u x u · x vx v · x u x v · x v⎡| |⎢= ⎣ x u x v| |⎤⎥⎦T ⎡| |⎢⎣ x u x v| |⎛⎡⎤⎞([]) x u · x u x u · x v 0EG − F 2 x u · x u x u · x v ⎜⎢⎥⎟= det= det ⎝⎣x v · x u x v · x v 0 ⎦⎠x v · x u x v · x v0 0 1⎛⎡⎤T ⎡ ⎤⎞⎛ ⎡ ⎤⎞| | | | | || | |= det ⎜⎢⎥ ⎢ ⎥⎝⎣x u x v n ⎦ ⎣ x u x v n ⎦⎟⎠ = ⎜ ⎢ ⎥⎟⎝det ⎣ x u x v n ⎦⎠| | | | | || | |which is the square of the volume of the parallelepiped spanned by x u , x v , <strong>and</strong> n. S<strong>in</strong>ce n is a unitvector orthogonal to the plane spanned by x u <strong>and</strong> x v , this is, <strong>in</strong> turn, the square of the area of theparallelogram spanned by x u <strong>and</strong> x v . That is,EG − F 2 = ‖x u × x v ‖ 2 ≠0.We rem<strong>in</strong>d the reader that we obta<strong>in</strong> the surface area of the parametrized surface x: U → M bycalculat<strong>in</strong>g the double <strong>in</strong>tegral∫∫ √‖x u × x v ‖dudv = EG − F 2 dudv.UU⎤⎥⎦ ,2,EXERCISES 2.11. Derive the formula given <strong>in</strong> Example 1(e) for the parametrization of the unit sphere.2. Compute I (i.e., E, F , <strong>and</strong> G) for the follow<strong>in</strong>g parametrized surfaces.*a. the sphere of radius a: x(u, v) =a(s<strong>in</strong> u cos v, s<strong>in</strong> u s<strong>in</strong> v, cos u)b. the torus: x(u, v) =((a + b cos u) cos v, (a + b cos u) s<strong>in</strong> v, b s<strong>in</strong> u) (0


§1. Parametrized <strong>Surfaces</strong> <strong>and</strong> the <strong>First</strong> Fundamental Form 41*4. Show that if all the normal l<strong>in</strong>es to a surface pass through a fixed po<strong>in</strong>t, then the surface is (aportion of) a sphere. (By the normal l<strong>in</strong>e to M at P we mean the l<strong>in</strong>e pass<strong>in</strong>g through P withdirection vector the unit normal at P .)5. Check that the parametrization x(u, v) isconformal if <strong>and</strong> only if E = G <strong>and</strong> F =0. (H<strong>in</strong>t:For =⇒, choose two convenient pairs of orthogonal directions.)*6. Check that the parametrization of the unit sphere by stereographic projection (see Example1(e)) is conformal.7. Consider the hyperboloid of one sheet, M, given by the equation x 2 + y 2 − z 2 =1.a. Show that x(u, v) =(cosh u cos v, cosh u s<strong>in</strong> v, s<strong>in</strong>h u) gives a parametrization of M as asurface of revolution.*b. F<strong>in</strong>d two parametrizations ( of M as a ruled surface α(u)+vβ(u).uv +1c. Show that x(u, v) =uv − 1 , u − vuv − 1 , u + v )gives a parametrization of M where bothuv − 1sets of parameter curves are rul<strong>in</strong>gs.♯ 8.Given a ruled surface x(u, v) =α(u)+vβ(u) with α ′ ≠0<strong>and</strong> ‖β‖ =1;suppose that α ′ (u),β(u), <strong>and</strong> β ′ (u) are l<strong>in</strong>early dependent for every u. Prove that locally one of the follow<strong>in</strong>g musthold:(i) β = const;(ii) there is a function λ so that α(u)+λ(u)β(u) =const;(iii) there is a function λ so that (α + λβ) ′ (u) isanonzero multiple of β(u) for every u.Describe the surface <strong>in</strong> each of these cases. (H<strong>in</strong>t: There are c 1 , c 2 , c 3 (functions of u), neversimultaneously 0, so that c 1 (u)α ′ (u)+c 2 (u)β(u)+c 3 (u)β ′ (u) =0 for all u. Consider separatelythe cases c 1 (u) =0<strong>and</strong> c 1 (u) ≠0.Inthe latter case, divide through.)9. (The Mercator projection) Mercator developed his system for mapp<strong>in</strong>g the earth, as pictured <strong>in</strong>uvFigure 1.8Figure 1.8, <strong>in</strong> 1569, about a century before the advent of calculus. We want a parametrizationx(u, v) ofthe sphere, u ∈ R, v ∈ [0, 2π), so that the u-curves are the longitudes <strong>and</strong> sothat the parametrization is conformal. Lett<strong>in</strong>g (φ, θ) bethe usual spherical coord<strong>in</strong>ates, write


42 Chapter 2. <strong>Surfaces</strong>: Local Theoryφ = f(u) <strong>and</strong> θ = v. Show that conformality <strong>and</strong> symmetry about the equator will dictatef(u) =2arctan(e −u ). Deduce that(Cf. Example 2 <strong>in</strong> Section 1 of Chapter 1.)x(u, v) =(sechu cos v, sechu s<strong>in</strong> v, tanh u).10. A parametrization x(u, v) iscalled a Tschebyschev net if the opposite sides of any quadrilateralformed by the coord<strong>in</strong>ate curves have equal length.a. Prove that this occurs if <strong>and</strong> only if ∂E∂v = ∂G =0. (H<strong>in</strong>t: Express the length of the∂uu-curves, u 0 ≤ u ≤ u 1 ,asan<strong>in</strong>tegral <strong>and</strong> use the fact that this length is <strong>in</strong>dependent ofv.)b. Prove that we can locally reparametrize by ˜x(ũ, ṽ) soastoobta<strong>in</strong> Ẽ = ˜G =1, ˜F =cos θ(ũ, ṽ) (so that the ũ- <strong>and</strong> ṽ-curves are parametrized by arclength <strong>and</strong> meet at angle/θ). (H<strong>in</strong>t: Choose ũ as a function of u so that ˜xũ = x u (dũ/du) has unit length.)11. Suppose x <strong>and</strong> y are two parametrizations of a surface M near P .Sayx(u 0 ,v 0 )=P = y(s 0 ,t 0 ).Prove that Span(x u , x v )=Span(y s , y t ) (where the partial derivatives are all evaluated at theobvious po<strong>in</strong>ts). (H<strong>in</strong>t: f = x −1 ◦y gives a C 1 map from an open set around (s 0 ,t 0 )toanopenset around (u 0 ,v 0 ). Apply the cha<strong>in</strong> rule to show y s , y t ∈ Span(x u , x v ).)12. (A programmable calculator or Maple may be useful for parts of this problem.) A catenoid, aspictured <strong>in</strong> Figure 1.9, is parametrized byx(u, v) =(a cosh u cos v, a cosh u s<strong>in</strong> v, au), u ∈ R, 0 ≤ v0 fixed).Figure 1.9*a. Compute the surface area of that portion of the catenoid given by |u| ≤ 1/a. (H<strong>in</strong>t:cosh 2 u = 1 2(1 + cosh 2u).)b. Given the pair of parallel circles x 2 + y 2 = R 2 , |z| =1,for what values of R is there atleast one catenoid with the circles as boundary? (H<strong>in</strong>t: Graph f(t) =t cosh(1/t).)c. For the values of R <strong>in</strong> part b, compare the area of the catenoid(s) with 2πR 2 , the area ofthe pair of disks fill<strong>in</strong>g <strong>in</strong> the circles. For what values of R does the pair of disks have theleast area?13. There are two obvious families of circles on a torus. F<strong>in</strong>d a third family. (H<strong>in</strong>t: Look for a planethat is tangent to the torus at two po<strong>in</strong>ts. Us<strong>in</strong>g the parametrization of the torus, you should


§2. The Gauss Map <strong>and</strong> the Second Fundamental Form 43be able to f<strong>in</strong>d parametric equations for the curve <strong>in</strong> which the bitangent plane <strong>in</strong>tersects thetorus.)2. The Gauss Map <strong>and</strong> the Second Fundamental FormGiven a regular parametrized surface M, the function n: M → Σ that assigns to each po<strong>in</strong>tP ∈ M the unit normal n(P ), as pictured <strong>in</strong> Figure 2.1, is called the Gauss map of M. Asweshallsee <strong>in</strong> this chapter, most of the geometric <strong>in</strong>formation about our surface M is encapsulated <strong>in</strong> thePnn(P)Figure 2.1mapp<strong>in</strong>g n.Example 1. A few basic examples are these.(a) On a plane, the tangent plane never changes, so the Gauss map is a constant.(b) On a cyl<strong>in</strong>der, the tangent plane is constant along the rul<strong>in</strong>gs, so the Gauss map sends theentire surface to an equator of the sphere.(c) On a sphere centered at the orig<strong>in</strong>, the Gauss map is merely the (normalized) positionvector.(d) On a saddle surface (as pictured <strong>in</strong> Figure 2.1), the Gauss map appears to “reverse orientation”:As we move counterclockwise <strong>in</strong> a small circle around P ,wesee that the unit vectorn turns clockwise around n(P ). ▽Recall from the Appendix that for any function f on M (scalar- or vector-valued) <strong>and</strong> anytangent vector V ∈ T P M,wecan compute the directional derivative D V f(P )bychoos<strong>in</strong>g a curveα: (−ε, ε) → M with α(0) = P <strong>and</strong> α ′ (0) = V <strong>and</strong> comput<strong>in</strong>g (f ◦α) ′ (0).To underst<strong>and</strong> the shape of M at the po<strong>in</strong>t P , might try to underst<strong>and</strong> the curvature at P ofvarious curves <strong>in</strong> M. Perhaps the most obvious th<strong>in</strong>g to try is various normal slices of M. That is,we slice M with the plane through P spanned by n(P ) <strong>and</strong> a unit vector V ∈ T P M.Various suchnormal slices are shown for a saddle surface <strong>in</strong> Figure 2.2. Let α be the arclength-parametrizedcurve obta<strong>in</strong>ed by tak<strong>in</strong>g such a normal slice. We have α(0) = P <strong>and</strong> α ′ (0) = V. Then s<strong>in</strong>ce thecurve lies the plane spanned by n(P ) <strong>and</strong> V, the pr<strong>in</strong>cipal normal of the curve at P must be ±n(P )(+ if the curve is curv<strong>in</strong>g towards n, − if it’s curv<strong>in</strong>g away). Apply<strong>in</strong>g Lemma 2.1 of Chapter 1


44 Chapter 2. <strong>Surfaces</strong>: Local TheoryFigure 2.2yet aga<strong>in</strong>, we have:(†)±κ = κN · n = T ′ (0) · n = −T · (n◦α) ′ (0) = −D V n(P ) · V.This leads us to study the directional derivative D V n(P ) more carefully.Proposition 2.1. For any V ∈ T P M, the directional derivative D V n(P ) ∈ T P M. Moreover,the l<strong>in</strong>ear map S P : T P M → T P M def<strong>in</strong>ed byS P (V) =−D V n(P )is a symmetric l<strong>in</strong>ear map; i.e., for any U, V ∈ T P M,wehave(∗)S P (U) · V = U · S P (V)S P is called the shape operator at P .Proof. Forany curve α: (−ε, ε) → M with α(0) = P <strong>and</strong> α ′ (0) = V, weobserve that n◦α hasconstant length 1. Thus, by Lemma 2.1 of Chapter 1, D V n(P )·n(P )=(n◦α) ′ (0)·(n◦α)(0) = 0, soD V n(P )is<strong>in</strong>the tangent plane to M at P . That S P is a l<strong>in</strong>ear map is an immediate consequenceof Proposition 2.3 of the Appendix.Symmetry is our first important application of the equality of mixed partial derivatives. <strong>First</strong>we verify (∗) when U = x u , V = x v . Note that n · x v =0,so0= ( n · x v)u = n u · x v + n · x vu .(Remember that we’re writ<strong>in</strong>g n u for D xu n.) Thus,S P (x u ) · x v = −D xu n(P ) · x v = −n u · x v = n · x vu= n · x uv = −n v · x u = −D xv n(P ) · x u = S P (x v ) · x u .


§2. The Gauss Map <strong>and</strong> the Second Fundamental Form 45Next, know<strong>in</strong>g this, we just write out general vectors U <strong>and</strong> V as l<strong>in</strong>ear comb<strong>in</strong>ations of x u <strong>and</strong>x v :IfU = ax u + bx v <strong>and</strong> V = cx u + dx v , thenS P (U) · V = S P (ax u + bx v ) · (cx u + dx v )= ( aS P (x u )+bS P (x v ) ) · (cx u + dx v )= acS P (x u ) · x u + adS P (x u ) · x v + bcS P (x v ) · x u + bdS P (x v ) · x v= acS P (x u ) · x u + adS P (x v ) · x u + bcS P (x u ) · x v + bdS P (x v ) · x v=(ax u + bx v ) · (cSP (x u )+dS P (x v ) ) = U · S P (V),as required.□Proposition 2.2. If the shape operator S P is O for all P ∈ M, then M is a subset of a plane.Proof. S<strong>in</strong>ce the directional derivative of the unit normal n is 0 <strong>in</strong> every direction at everypo<strong>in</strong>t P ,wehaven u = n v = 0 for any (local) parametrization x(u, v) ofM. ByProposition 2.4 ofthe Appendix, it follows that n is constant. (This is why we assume our surfaces are connected.) □Example 2. Let M be a sphere of radius a centered at the orig<strong>in</strong>. Then n = 1 x(u, v), so foraany P ,wehaveS P (x u )=−n u = − 1 a x u <strong>and</strong> S P (x v )=−n v = − 1 a x v,soS Pidentity map on each tangent plane. ▽is −1/a times theIt does not seem an easy task to give the matrix of the shape operator with respect to thebasis {x u , x v }. But, <strong>in</strong> general, the proof of Proposition 2.1 suggests that we def<strong>in</strong>e the secondfundamental form, as follows. If U, V ∈ T P M,wesetII P (U, V) =S P (U) · V.Note that the formula (†) onp.44shows that the curvature of the normal slice <strong>in</strong> direction V (with‖V‖ =1)is, <strong>in</strong> our new notation, given by±κ = −D V n(P ) · V = S P (V) · V =II P (V, V).As we did at the end of the previous section, we wish to give a matrix representation whenwe’re work<strong>in</strong>g with a parametrized surface. As we saw <strong>in</strong> the proof of Proposition 2.1, we havel =II P (x u , x u )= −D xu n · x u = x uu · nm =II P (x u , x v )= −D xu n · x v = x vu · n = x uv · n =II P (x v , x u )n =II P (x v , x v )=−D xv n · x v = x vv · n.(By the way, this expla<strong>in</strong>s the presence of the m<strong>in</strong>us sign <strong>in</strong> the orig<strong>in</strong>al def<strong>in</strong>ition of the shapeoperator.) We then write[ ] []l m x uu · n x uv · nII P = =.m n x uv · n x vv · nIf, as before, U = ax u + bx v <strong>and</strong> V = cx u + dx v , thenII P (U, V) =II P (ax u + bx v ,cx u + dx v )= acII P (x u , x u )+adII P (x u , x v )+bcII P (x v , x u )+bdII P (x v , x v )


46 Chapter 2. <strong>Surfaces</strong>: Local Theory= l(ac)+m(bc + ad)+n(bd).In the event that {x u , x v } is an orthonormal basis for T P M,wesee that the matrix II P representsthe shape operator S P . But it is not difficult to check (see Exercise 2) that, <strong>in</strong> general, the matrixof the l<strong>in</strong>ear map S P with respect to the basis {x u , x v } is given byI −1P II P =[EF] −1 [F lG m]m.nRemark. We proved <strong>in</strong> Proposition 2.1 that S P is a symmetric l<strong>in</strong>ear map. This means thatits matrix representation with respect to an orthonormal basis (or, more generally, orthogonal basiswith vectors of equal length) will be symmetric: In this case the matrix I P is a scalar multiple ofthe identity matrix <strong>and</strong> the matrix product rema<strong>in</strong>s symmetric.By the Spectral Theorem, Theorem 1.3 of the Appendix, S P has two real eigenvalues, traditionallydenoted k 1 (P ), k 2 (P ).Def<strong>in</strong>ition. The eigenvalues of S P are called the pr<strong>in</strong>cipal curvatures of M at P . Correspond<strong>in</strong>geigenvectors are called pr<strong>in</strong>cipal directions. A curve <strong>in</strong> M is called a l<strong>in</strong>e of curvature if itstangent vector at each po<strong>in</strong>t is a pr<strong>in</strong>cipal direction.Recall that it also follows from the Spectral Theorem that the pr<strong>in</strong>cipal directions are orthogonal,so we can always choose an orthonormal basis for T P M consist<strong>in</strong>g of pr<strong>in</strong>cipal directions. Hav<strong>in</strong>gdone so, we can then easily determ<strong>in</strong>e the curvatures of normal slices <strong>in</strong> arbitrary directions, asfollows.Proposition 2.3 (Euler’s Formula). Let e 1 , e 2 be unit vectors <strong>in</strong> the pr<strong>in</strong>cipal directions atP with correspond<strong>in</strong>g pr<strong>in</strong>cipal curvatures k 1 <strong>and</strong> k 2 . Suppose V = cos θe 1 + s<strong>in</strong> θe 2 for someθ ∈ [0, 2π), aspictured <strong>in</strong> Figure 2.3. Then II P (V, V) =k 1 cos 2 θ + k 2 s<strong>in</strong> 2 θ.e 2Vθ e 1Figure 2.3Proof. This is a straightforward computation: S<strong>in</strong>ce S P (e i )=k i e i for i =1, 2, we haveII P (V, V) =S P (V) · V = S P (cos θe 1 + s<strong>in</strong> θe 2 ) · (cos θe 1 + s<strong>in</strong> θe 2 )= (cos θk 1 e 1 + s<strong>in</strong> θk 2 e 2 ) · (cos θe 1 + s<strong>in</strong> θe 2 )=k 1 cos 2 θ + k 2 s<strong>in</strong> 2 θ,as required.□


§2. The Gauss Map <strong>and</strong> the Second Fundamental Form 47On a sphere, all normal slices have the same (nonzero) curvature. On the other h<strong>and</strong>, if we lookcarefully at Figure 2.2, we see that certa<strong>in</strong> normal slices of a saddle surface are true l<strong>in</strong>es. Thisleads us to make the follow<strong>in</strong>gDef<strong>in</strong>ition. If the normal slice <strong>in</strong> direction V has zero curvature, i.e., if II P (V, V) =0, thenwe call V an asymptotic direction. 2 A curve <strong>in</strong> M is called an asymptotic curve if its tangent vectorat each po<strong>in</strong>t is an asymptotic direction.Example 3. If a surface M conta<strong>in</strong>s a l<strong>in</strong>e, that l<strong>in</strong>e is an asymptotic curve. For the normalslice <strong>in</strong> the direction of the l<strong>in</strong>e conta<strong>in</strong>s the l<strong>in</strong>e (<strong>and</strong> perhaps other th<strong>in</strong>gs far away), which, ofcourse, has zero curvature. ▽Corollary 2.4. There is an asymptotic direction at P if <strong>and</strong> only if k 1 k 2 ≤ 0.Proof. k 2 =0if <strong>and</strong> only if e 2 is an asymptotic direction. Now suppose k 2 ≠0. IfV isa unit asymptotic vector mak<strong>in</strong>g angle θ with e 1 , then we have k 1 cos 2 θ + k 2 s<strong>in</strong> 2 θ =0,<strong>and</strong> sotan 2 θ = −k 1 /k 2 ,sok 1 k 2 ≤ 0. Conversely, if k 1 k 2 ≤ 0, take θ with tan θ = ± √ −k 1 /k 2 , <strong>and</strong> thenV is an asymptotic direction. □Example 4. We consider the helicoid, as pictured <strong>in</strong> Figure 1.2. It is a ruled surface <strong>and</strong> sothe rul<strong>in</strong>gs are asymptotic curves. What is quite less obvious is that the family of helices on thesurface are also asymptotic curves. But, as we see <strong>in</strong> Figure 2.4, the normal slice tangent to thePFigure 2.4helix at P has an <strong>in</strong>flection po<strong>in</strong>t at P , <strong>and</strong> therefore the helix is an asymptotic curve. We ask thereader to check this by calculation <strong>in</strong> Exercise 5. ▽It is also an immediate consequence of Proposition 2.3 that the pr<strong>in</strong>cipal curvatures are themaximum <strong>and</strong> m<strong>in</strong>imum (signed) curvatures of the various normal slices. Assume k 2 ≤ k 1 . Thenk 1 cos 2 θ + k 2 s<strong>in</strong> 2 θ = k 1 (1 − s<strong>in</strong> 2 θ)+k 2 s<strong>in</strong> 2 θ = k 1 +(k 2 − k 1 ) s<strong>in</strong> 2 θ ≤ k 12 Of course, V ̸=0 here. See Exercise 21 for an explanation of this term<strong>in</strong>ology.


48 Chapter 2. <strong>Surfaces</strong>: Local Theory(<strong>and</strong>, similarly, ≥ k 2 ). Moreover, as the Spectral Theorem tells us, the maximum <strong>and</strong> m<strong>in</strong>imumoccur at right angles to one another. Look<strong>in</strong>g back at Figure 2.2, where the slices are taken at angles<strong>in</strong> <strong>in</strong>crements of π/8, we see that the normal slices that are “most curved” appear <strong>in</strong> the third <strong>and</strong>seventh frames; the asymptotic directions appear <strong>in</strong> the second <strong>and</strong> fourth frames. (Cf. Exercise8.)Next we come to one of the most important concepts <strong>in</strong> the geometry of surfaces:Def<strong>in</strong>ition. The product of the pr<strong>in</strong>cipal curvatures is called the Gaussian curvature: K =det S P = k 1 k 2 . The average of the pr<strong>in</strong>cipal curvatures is called the mean curvature: H = 1 2 trS P =12 (k 1 + k 2 ). We say M is a m<strong>in</strong>imal surface if H =0.Note that whereas the signs of the pr<strong>in</strong>cipal curvatures change if we reverse the direction of theunit normal n, the Gaussian curvature K, be<strong>in</strong>g the product of both, is <strong>in</strong>dependent of the choiceof unit normal. (And the sign of the mean curvature depends on the choice.)Def<strong>in</strong>ition. Fix P ∈ M. We say P is an umbilic 3 if k 1 = k 2 . If k 1 = k 2 =0,wesayP is aplanar po<strong>in</strong>t. IfK =0but P is not a planar po<strong>in</strong>t, we say P is a parabolic po<strong>in</strong>t. IfK>0, we sayP is an elliptic po<strong>in</strong>t, <strong>and</strong> if K


§2. The Gauss Map <strong>and</strong> the Second Fundamental Form 49rectangle collapses to a l<strong>in</strong>e segment; for a saddle surface, orientation is reversed by n <strong>and</strong> so theGaussian curvature is negative.)Let’s close this section by revisit<strong>in</strong>g our discussion of the curvature of normal slices. Suppose αis an arclength-parametrized curve ly<strong>in</strong>g on M with α(0) = P <strong>and</strong> α ′ (0) = V. Then the calculation<strong>in</strong> formula (†) onp.44shows thatII P (V, V) =κN · n;i.e., II P (V, V) gives the component of the curvature vector κN of α normal to the surface M atP , which we denote by κ n <strong>and</strong> call the normal curvature of α at P . What is remarkable aboutthis formula is that it shows that the normal curvature depends only on the direction of α at P<strong>and</strong> otherwise not on the curve. (For the case of the normal slice, the normal curvature is, up to asign, all the curvature.) We immediately deduce the follow<strong>in</strong>gProposition 2.5 (Meusnier’s Formula). Let α be acurve on M pass<strong>in</strong>g through P with unittangent vector V. ThenII P (V, V) =κ n = κ cos φ,where φ is the angle between the pr<strong>in</strong>cipal normal, N, ofα <strong>and</strong> the surface normal, n, atP .In particular, if α is an asymptotic curve, then its normal curvature is 0 at each po<strong>in</strong>t.Example 6. Let’s now <strong>in</strong>vestigate a very <strong>in</strong>terest<strong>in</strong>g surface, called the pseudosphere, asshown<strong>in</strong> Figure 2.6. It is the surface of revolution obta<strong>in</strong>ed by rotat<strong>in</strong>g the tractrix (see Example 2 ofnNFigure 2.6Chapter 1, Section 1) about the x-axis, <strong>and</strong> so it is parametrized byx(u, v) =(u − tanh u, sechu cos v, sechu s<strong>in</strong> v),u > 0, v∈ [0, 2π).Note that the circles (of revolution) are l<strong>in</strong>es of curvature: Either apply Exercise 13 or observe,directly, that the only component of the surface normal that changes as we move around the circle isnormal to the circle <strong>in</strong> the plane of the circle. Similarly, the various tractrices are l<strong>in</strong>es of curvature:In the plane of one tractrix, the surface normal <strong>and</strong> the curve normal agree.Now, by Exercise 1.2.5, the curvature of the tractrix is κ =1/ s<strong>in</strong>h u; s<strong>in</strong>ce N = n along thiscurve, we have k 1 = κ n =1/ s<strong>in</strong>h u. Now what about the circles? Here we have κ =1/sechu =cosh u, but this is not the normal curvature. The angle θ we see <strong>in</strong> Figure 1.9 of Chapter 1 isthe same as the angle φ between N <strong>and</strong> n <strong>in</strong> Figure 2.6 (to see why, see Figure 2.7). Thus, by


50 Chapter 2. <strong>Surfaces</strong>: Local TheorynN1tanh uθFigure 2.7Meusnier’s Formula, Proposition 2.5, we have k 2 = κ n = κ cos φ = (cosh u)(− tanh u) =− s<strong>in</strong>h u.Amaz<strong>in</strong>gly, then, we have K = k 1 k 2 =(1/ s<strong>in</strong>h u)(− s<strong>in</strong>h u) =−1. ▽Example 7. Let’s now consider the case of a general surface of revolution, parametrized as <strong>in</strong>Example 2 of Section 1, byx(u, v) = ( f(u) cos v, f(u) s<strong>in</strong> v, g(u) ) ,where f ′ (u) 2 +g ′ (u) 2 =1. Recall that the u-curves are called meridians <strong>and</strong> the v-curves are calledparallels. Then<strong>and</strong> so we havex u = ( f ′ (u) cos v, f ′ (u) s<strong>in</strong> v, g ′ (u) )x v = ( −f(u) s<strong>in</strong> v, f(u) cos v, 0 )n = ( −g ′ (u) cos v, −g ′ (u) s<strong>in</strong> v, f ′ (u) )x uu = ( f ′′ (u) cos v, f ′′ (u) s<strong>in</strong> v, g ′′ (u) )x uv = ( −f ′ (u) s<strong>in</strong> v, f ′ (u) cos v, 0 )x vv = ( −f(u) cos v, −f(u) s<strong>in</strong> v, 0 ) ,E =1, F =0, G = f(u) 2 <strong>and</strong> l = f ′ (u)g ′′ (u) − f ′′ (u)g ′ (u), m =0, n = f(u)g ′ (u).By Exercise 2.2.1, then k 1 = f ′ (u)g ′′ (u) − f ′′ (u)g ′ (u) <strong>and</strong> k 2 = g ′ (u)/f(u). Thus,K = k 1 k 2 = ( f ′ (u)g ′′ (u) − f ′′ (u)g ′ (u) ) g ′ (u)f(u) = −f′′ (u)f(u) ,s<strong>in</strong>ce from f ′ (u) 2 + g ′ (u) 2 =1we deduce that f ′ (u)f ′′ (u)+g ′ (u)g ′′ (u) =0,<strong>and</strong> sof ′ (u)g ′ (u)g ′′ (u) − f ′′ (u)g ′ (u) 2 = −(f ′ (u) 2 + g ′ (u) 2 )f ′′ (u) =−f ′′ (u).Note, as we observed <strong>in</strong> the special case of Example 6, that on every surface of revolution, themeridians <strong>and</strong> the parallels are l<strong>in</strong>es of curvature. ▽


§2. The Gauss Map <strong>and</strong> the Second Fundamental Form 51EXERCISES 2.2*1. Check that the parameter curves are l<strong>in</strong>es of curvature if <strong>and</strong> only if F = m =0. Show,moreover, that <strong>in</strong> this event, we have the pr<strong>in</strong>cipal curvatures k 1 = l/E <strong>and</strong> k 2 = n/G.♯ 2. a. Show that the matrix represent<strong>in</strong>g the l<strong>in</strong>ear map S P : T P M → T P M with respect to thebasis {x u , x v } isI −1P II P =[EF] −1 [F lG m]m.n(H<strong>in</strong>t: It suffices to check S P (x u ) · x u , S P (x u ) · x v , <strong>and</strong> so on.)ln − m2b. Deduce that K =EG − F 2 .3. Compute the second fundamental form II P of the follow<strong>in</strong>g parametrized surfaces. Then calculatethe matrix of the shape operator, determ<strong>in</strong>e H <strong>and</strong> K.a. the cyl<strong>in</strong>der: x(u, v) =(a cos u, a s<strong>in</strong> u, v)*b. the torus: x(u, v) =((a + b cos u) cos v, (a + b cos u) s<strong>in</strong> v, b s<strong>in</strong> u) (0


52 Chapter 2. <strong>Surfaces</strong>: Local Theoryc. (More challeng<strong>in</strong>g) Show that, more generally, for any θ <strong>and</strong> m ≥ 3, we haveH = 1 (( 2π ) ( 2π(m − 1) ) )κ n (θ)+κ n θ + + ···+ κn θ + .mmm10. a. Apply Meusnier’s Formula to a latitude circle on a sphere of radius a to calculate thenormal curvature.b. Prove that the curvature of any curve ly<strong>in</strong>g on the sphere of radius a satisfies κ ≥ 1/a.11. Prove or give a counterexample: If M is a surface with Gaussian curvature K>0, then thecurvature of any curve C ⊂ M is everywhere positive. (Remember that, by def<strong>in</strong>ition, κ ≥ 0.)♯ 12.Suppose that for every P ∈ M, the shape operator S P is some scalar multiple of the identity,i.e., S P (V) =k(P )V for all V ∈ T P M. (Here the scalar k(P )maywell depend on the po<strong>in</strong>tP .)a. Differentiate the equationsD xu n = n u = − kx uD xv n = n v = − kx vappropriately to determ<strong>in</strong>e k u <strong>and</strong> k v <strong>and</strong> deduce that k must be constant.b. We showed <strong>in</strong> Proposition 2.2 that M is planar when k =0. Show that when k ≠0,M is(a portion of) a sphere.13. a. Prove that if α is a l<strong>in</strong>e of curvature <strong>in</strong> M, then (n◦α) ′ (t) =−k(t)α ′ (t), where k(t) isthepr<strong>in</strong>cipal curvature at α(t) <strong>in</strong>the direction α ′ (t). (More colloquially, differentiat<strong>in</strong>g alongthe curve α, wejust write n ′ = −kα ′ .)b. Suppose two surfaces M <strong>and</strong> M ∗ <strong>in</strong>tersect along a curve C. Suppose C is a l<strong>in</strong>e of curvature<strong>in</strong> M. Prove that C is a l<strong>in</strong>e of curvature <strong>in</strong> M ∗ if <strong>and</strong> only if the angle between M <strong>and</strong>M ∗ is constant along C. (In the proof of ⇐=, be sure to <strong>in</strong>clude the case that M <strong>and</strong> M ∗<strong>in</strong>tersect tangentially along C.)14. Prove or give a counterexample:a. If a curve is both an asymptotic curve <strong>and</strong> a l<strong>in</strong>e of curvature, then it must be planar.b. If a curve is planar <strong>and</strong> an asymptotic curve, then it must be a l<strong>in</strong>e.15. a. How is the Frenet frame along an asymptotic curve related to the geometry of the surface?b. Suppose K(P ) < 0. If C is an asymptotic curve with κ(P ) ≠0,prove that its torsionsatisfies |τ(P )| = √ −K(P ). (H<strong>in</strong>t: If we choose an orthonormal basis {U, V} for T P (M)with U tangent to C, what is the matrix for S P ? See the Remark on p. 46.)16. Cont<strong>in</strong>u<strong>in</strong>g Exercise 15, show that if K(P ) < 0, then the two asymptotic curves have torsionof opposite signs at P .17. Prove that the only m<strong>in</strong>imal ruled surface with no planar po<strong>in</strong>ts is the helicoid. (H<strong>in</strong>t: Considerthe curves orthogonal to the rul<strong>in</strong>gs. Use Exercises 8b <strong>and</strong> 1.2.19.)


§2. The Gauss Map <strong>and</strong> the Second Fundamental Form 5318. Suppose U ⊂ R 3 is open <strong>and</strong> x: U → R 3 is a smooth map (of rank 3) so that x u , x v , <strong>and</strong> x ware always orthogonal. Then the level surfaces u = const, v = const, w = const form a triplyorthogonal system of surfaces.a. Show that the spherical coord<strong>in</strong>ate mapp<strong>in</strong>g x(u, v, w) =(u s<strong>in</strong> v cos w, u s<strong>in</strong> v s<strong>in</strong> w, u cos v)furnishes an example.b. Prove that the curves of <strong>in</strong>tersection of any pair of surfaces from different systems (e.g.,v = const <strong>and</strong> w = const) are l<strong>in</strong>es of curvature <strong>in</strong> each of the respective surfaces. (H<strong>in</strong>t:Differentiate the various equations x u · x v =0,x v · x w =0,x u · x w =0with respect to themiss<strong>in</strong>g variable. What are the shape operators of the various surfaces?)19. In this exercise we analyze the surfaces of revolution that are m<strong>in</strong>imal. It will be convenient towork with a meridian as a graph (y = h(u), z = u), as opposed to our customary parametrizationof surfaces of revolution.a. Use Exercise 1.2.4 <strong>and</strong> Proposition 2.5 to show that the pr<strong>in</strong>cipal curvatures arek 1 =h ′′(1 + h ′2 ) 3/2 <strong>and</strong> k 2 = − 1 h · 1√ .1+h ′2b. Deduce that H =0if <strong>and</strong> only if h(u)h ′′ (u) =1+h ′ (u) 2 .c. Solve the differential equation. (H<strong>in</strong>t: Either substitute z(u) =lnh(u)or<strong>in</strong>troduce w(u) =h ′ (u), f<strong>in</strong>d dw/dh, <strong>and</strong> <strong>in</strong>tegrate by separat<strong>in</strong>g variables.) You should f<strong>in</strong>d that h(u) =1ccosh(cu + b) for some constants b <strong>and</strong> c.20. By choos<strong>in</strong>g coord<strong>in</strong>ates <strong>in</strong> R 3 appropriately, we may arrange that P is the orig<strong>in</strong>, the tangentplane T P M is the xy-plane, <strong>and</strong> the x- <strong>and</strong> y-axes are <strong>in</strong> the pr<strong>in</strong>cipal directions at P .a. Show that <strong>in</strong> these coord<strong>in</strong>ates M is locally the graph z = f(x, y) = 1 2 (k 1x 2 + k 2 y 2 )+...,...where limx,y→0 x 2 + y 2 =0.(Youmay start with Taylor’s Theorem: if f is C2 ,wehavewheref(x, y) =f(0, 0) + f x (0, 0)x + f y (0, 0)y +limx,y→0...x 2 + y 2 = 0.)12(fxx (0, 0)x 2 +2f xy (0, 0)xy + f yy (0, 0)y 2) + ...,b. Show that if P is an elliptic po<strong>in</strong>t, then a neighborhood of P <strong>in</strong> M ∩ T P M is just theorig<strong>in</strong> itself. Show that if P is a hyperbolic po<strong>in</strong>t, such a neighborhood is a curve thatcrosses itself at P <strong>and</strong> whose tangent directions at P are the asymptotic directions. Whathappens <strong>in</strong> the case of a parabolic po<strong>in</strong>t?21. Let P ∈ M be a non-planar po<strong>in</strong>t, <strong>and</strong> if K ≥ 0, choose the unit normal so that l, n ≥ 0.a. We def<strong>in</strong>e the Dup<strong>in</strong> <strong>in</strong>dicatrix to be the conic <strong>in</strong> T P M def<strong>in</strong>ed by the equation II P (V, V) =1. Show that if P is an elliptic po<strong>in</strong>t, the Dup<strong>in</strong> <strong>in</strong>dicatrix is an ellipse; if P is a hyperbolicpo<strong>in</strong>t, the Dup<strong>in</strong> <strong>in</strong>dicatrix is a hyperbola; <strong>and</strong> if P is a parabolic po<strong>in</strong>t, the Dup<strong>in</strong><strong>in</strong>dicatrix is a pair of parallel l<strong>in</strong>es.b. Show that if P is a hyperbolic po<strong>in</strong>t, the asymptotes of the Dup<strong>in</strong> <strong>in</strong>dicatrix are given byII P (V, V) =0, i.e., the set of asymptotic directions.


54 Chapter 2. <strong>Surfaces</strong>: Local Theoryc. Suppose M is represented locally near P as <strong>in</strong> Exercise 20. Show that for small positivevalues of c, the <strong>in</strong>tersection of M with the plane z = c “looks like” the Dup<strong>in</strong> <strong>in</strong>dicatrix.How can you make this statement more precise?22. Suppose the surface M is given near P as a level surface of a smooth function f : R 3 → R with∇f(P ) ≠0.Al<strong>in</strong>e L ⊂ R 3 is said to have (at least) k-po<strong>in</strong>t contact with M at P if, given anyl<strong>in</strong>ear parametrization α of L with α(0) = P , the function f = f ◦α vanishes to order k − 1,i.e., f(0) = f ′ (0) = ···= f (k−1) (0) = 0. (Such a l<strong>in</strong>e is to be visualized as the limit of l<strong>in</strong>es that<strong>in</strong>tersect M at P <strong>and</strong> at k − 1 other po<strong>in</strong>ts that approach P .)a. Show that L has 2-po<strong>in</strong>t contact with M at P if <strong>and</strong> only if L is tangent to M at P , i.e.,L ⊂ T P M.b. Show that L has 3-po<strong>in</strong>t contact with M at P if <strong>and</strong> only if L is an asymptotic directionat P . (H<strong>in</strong>t: It may be helpful to follow the setup of Exercise 20.)c. (Challenge) What does it mean for L to have 4-po<strong>in</strong>t contact with M at P ?3. The Codazzi <strong>and</strong> Gauss Equations <strong>and</strong> the Fundamental Theorem of SurfaceTheoryWe now wish to proceed towards a deeper underst<strong>and</strong><strong>in</strong>g of Gaussian curvature. We have tothis po<strong>in</strong>t considered only the normal components of the second derivatives x uu , x uv , <strong>and</strong> x vv .Nowlet’s consider them <strong>in</strong> toto. S<strong>in</strong>ce {x u , x v , n} gives a basis for R 3 , there are functions Γuu, u Γuu,vΓuv u =Γvu, u Γuv v =Γvu, v Γvv, u <strong>and</strong> Γvv v so that(†)x uu =Γ u uux u +Γ v uux v + lnx uv =Γ u uvx u +Γ v uvx v + mnx vv =Γ uvvx u +Γ vvvx v + nn.(Note that x uv = x vu dictates the symmetries Γ uv • =Γ vu.) • The functions Γ •• • are called Christoffelsymbols.Example 1. Let’s compute the Christoffel symbols for the usual parametrization of the sphere.By straightforward calculation we obta<strong>in</strong>x u = (cos u cos v, cos u s<strong>in</strong> v, − s<strong>in</strong> u)x v =(− s<strong>in</strong> u s<strong>in</strong> v, s<strong>in</strong> u cos v, 0)x uu =(− s<strong>in</strong> u cos v, − s<strong>in</strong> u s<strong>in</strong> v, − cos u) =−x(u, v)x uv =(− cos u s<strong>in</strong> v, cos u cos v, 0)x vv =(− s<strong>in</strong> u cos v, − s<strong>in</strong> u s<strong>in</strong> v, 0) = − s<strong>in</strong> u(cos v, s<strong>in</strong> v, 0).(Note that the u-curves are great circles, parametrized by arclength, so it is no surprise that theacceleration vector x uu is <strong>in</strong>ward-po<strong>in</strong>t<strong>in</strong>g of length 1. The v-curves are latitude circles of radiuss<strong>in</strong> u, so, similarly, the acceleration vector x vv po<strong>in</strong>ts <strong>in</strong>wards towards the center of the respectivecircle.)


§3. The Codazzi <strong>and</strong> Gauss Equations <strong>and</strong> the Fundamental Theorem of Surface Theory 55S<strong>in</strong>ce x uu lies entirely <strong>in</strong> the direction of n, wehave Γuu u =Γuu v =0. Now, by<strong>in</strong>spection,x uv = cot ux v ,soΓuv u =0<strong>and</strong> Γuv v = cot u. Last, as we can see <strong>in</strong> Figure 3.1, we have x vv =nx vvs<strong>in</strong> u cos uus<strong>in</strong> uux uFigure 3.1− s<strong>in</strong> u cos ux u − s<strong>in</strong> 2 un, soΓ uvv = − s<strong>in</strong> u cos u <strong>and</strong> Γ vvv =0.▽Now, dott<strong>in</strong>g the equations <strong>in</strong> (†) with x u <strong>and</strong> x v givesx uu · x u =ΓuuE u +ΓuuFvx uu · x v =ΓuuF u +ΓuuGvx uv · x u =ΓuvE u +ΓuvFvx uv · x v =ΓuvF u +ΓuvGvNow observe thatx vv · x u =Γ uvvE +Γ vvvFx vv · x v =Γ uvvF +Γ vvvG.x uu · x u = 1 2 (x u · x u ) u = 1 2 E ux uv · x u = 1 2 (x u · x u ) v = 1 2 E vx uv · x v = 1 2 (x v · x v ) u = 1 2 G ux uu · x v =(x u · x v ) u − x u · x uv = F u − 1 2 E vx vv · x u =(x u · x v ) v − x uv · x v = F v − 1 2 G ux vv · x v = 1 2 (x v · x v ) v = 1 2 G vThus, we can rewrite our equations as follows:[ ][ ] [ ]E F Γuuu 1F G Γuuv =2 E uF u − 1 2 E =⇒v[ ][ ] [ ]E F Γuvu 1F G Γuvv =2 E v(‡)12 G =⇒u[ ][ ] [ ]E F Γvvu F v − 1F G Γvvv =2 G u12 G =⇒v[Γ u uuΓ v uu[Γ u uvΓ v uv[Γ uvvΓ vvv]=[E=F] [E=F] [EF] −1 [ ]1F2 E uG F u − 1 2 E v] −1 [ ]F 12 E v1G2 G u] −1 [ ]F F v − 1 2 G u1G2 G .v


56 Chapter 2. <strong>Surfaces</strong>: Local TheoryWhat is quite remarkable about these formulas is that the Christoffel symbols, which tell us aboutthe tangential component of the second derivatives x •• , can be computed just from know<strong>in</strong>g E, F ,<strong>and</strong> G, i.e., the first fundamental form.Example 2. Let’s now recompute the Christoffel symbols of the unit sphere <strong>and</strong> compare ouranswers with Example 1. S<strong>in</strong>ce E =1,F =0,<strong>and</strong> G = s<strong>in</strong> 2 u,wehave[ ] [ ][ ] [ ]Γuuu 1 0 0 0Γuuv =0 csc 2 =u 0 0[ ] [ ][ ] [ ]Γuvu 1 0 00Γuvv =0 csc 2 =u s<strong>in</strong> u cos u cot u[ ] [ ][] []Γvvu 1 0 − s<strong>in</strong> u cos u − s<strong>in</strong> u cos uΓvvv =0 csc 2 =.u 00Thus, the only nonzero Christoffel symbols are Γuv v =Γvu v = cot u <strong>and</strong> Γvv u = − s<strong>in</strong> u cos u, asbefore.▽By Exercise 2.2.2, the matrix of the shape operator S P with respect to the basis {x u , x v } is[ ] [ ] −1 [ ][]a c E F l m 1 lG − mF mG − nF==b d F G m n EG − F 2 .−lF + mE −mF + nENote that these coefficients tell us the derivatives of n with respect to u <strong>and</strong> v:(††)n u = D xu n = −S P (x u )=−(ax u + bx v )n v = D xv n = −S P (x v )=−(cx u + dx v ).We now differentiate the equations (†) aga<strong>in</strong> <strong>and</strong> use equality of mixed partial derivatives. Tostart, we havex uuv =(Γuu) u v x u +Γuux u uv +(Γuu) v v x v +Γuux v vv + l v n + ln v=(Γuu) u v x u +Γuu( u Γuuv x u +Γuvx v v + mn ) +(Γuu) v v x v +Γuuv (Γuvv x u +Γvvx v v + nn)+ l v n − l(cx u + dx v )= ( (Γuu) u v +ΓuuΓ u uv u +ΓuuΓ v vv u − lc ) x u + ( (Γuu) v v +ΓuuΓ u uv v +ΓuuΓ v vv v − ld ) x v+ ( Γuum u +Γuun v )+ l v n,<strong>and</strong>, similarly,x uvu = ( (Γ u uv) u +Γ u uvΓ u uu +Γ v uvΓ u uv − ma ) x u + ( (Γ v uv) u +Γ u uvΓ v uu +Γ v uvΓ v uv − mb ) x v+ ( lΓ u uv + mΓ v uv + m u)n.S<strong>in</strong>ce x uuv = x uvu ,wecompare the <strong>in</strong>dicated components <strong>and</strong> obta<strong>in</strong>:(x u ): (Γuu) u v +ΓuuΓ v vv u − lc =(Γuv) u u +ΓuvΓ v uv u − ma(♦) (x v ): (Γuu) v v +ΓuuΓ u uv v +ΓuuΓ v vv v − ld =(Γuv) v u +ΓuvΓ u uu v +ΓuvΓ v uv v − mb(n): l v + mΓuu u + nΓuu v = m u + lΓuv u + mΓuv.v


§3. The Codazzi <strong>and</strong> Gauss Equations <strong>and</strong> the Fundamental Theorem of Surface Theory 57Analogously, compar<strong>in</strong>g the <strong>in</strong>dicated components of x uvv = x vvu ,wef<strong>in</strong>d:(x u ): (Γuv) u v +ΓuvΓ u uv u +ΓuvΓ v vv u − mc =(Γvv) u u +Γ u(x v ): (Γuv) v v +ΓuvΓ u uv v − md =(Γvv) v u +ΓvvΓ uuu v − nb(n): m v + mΓuv u + nΓuv v = n u + lΓvv u + mΓvv.vThe two equations com<strong>in</strong>g from the normal component give us theCodazzi equationsvvΓ u uu +Γ vvvΓ u uv − nal v − m u = lΓ u uv + m ( Γ v uv − Γ u uu)− nΓvuum v − n u = lΓ uvv + m ( Γ vvv − Γ u uv)− nΓvuv .ln − m2Us<strong>in</strong>g K =EG − F 2 <strong>and</strong> the formulas above for a, b, c, <strong>and</strong> d, the four equations <strong>in</strong>volv<strong>in</strong>g the x u<strong>and</strong> x v components yield theGauss equationsEK = ( Γuuv )v − ( Γuvv )u +Γu uuΓuv v +ΓuuΓ v vFK = ( Γuvu )u − ( Γuuu )v +Γv uvΓuv u − ΓuuΓ v vvuFK = ( Γuvv )v − ( Γvvv )u +Γu uvΓuv v − ΓvvΓ uuuvGK = ( Γ uvv)u − ( Γ u uv)v +Γu vvΓ u uu +Γ vvvΓ u uv − ( Γ u uvFor example, to derive the first, we use the equation (♦) above:(Γvuu)v − ( Γ v uvvv − Γ u uvΓ v uu − ( Γ v uv)u +Γu uuΓ v uv +Γ v uuΓ vvv − Γ u uvΓ v uu − ( Γ v uv) 2 = ld − mb=) 2) 2 − Γvuv Γ uvv.1 ( ) E(ln − m 2 )l(−mF + nE)+m(lF − mE) =EG − F 2 EG − F 2 = EK.In an orthogonal parametrization (F = 0), we leave it to the reader to check <strong>in</strong> Exercise 2 thatK = − 1 ( (2 √ Ev) (√EG+ Gu) )(∗)√ .EG v EG uOne of the crown<strong>in</strong>g results of local differential geometry is the follow<strong>in</strong>gTheorem 3.1 (Gauss’s Theorema Egregium). The Gaussian curvature is determ<strong>in</strong>ed by onlythe first fundamental form I. That is, K can be computed from just E, F , G, <strong>and</strong> their first <strong>and</strong>second partial derivatives.Proof. From any of the Gauss equations, we see that K can be computed by know<strong>in</strong>g any oneof E, F , <strong>and</strong> G, together with the Christoffel symbols <strong>and</strong> their derivatives. But the equations (‡)show that the Christoffel symbols (<strong>and</strong> hence any of their derivatives) can be calculated <strong>in</strong> termsof E, F , <strong>and</strong> G <strong>and</strong> their partial derivatives. □Corollary 3.2. If two surfaces are locally isometric, their Gaussian curvatures at correspond<strong>in</strong>gpo<strong>in</strong>ts are equal.


58 Chapter 2. <strong>Surfaces</strong>: Local TheoryFor example, the plane <strong>and</strong> cyl<strong>in</strong>der are locally isometric, <strong>and</strong> hence the cyl<strong>in</strong>der (as we wellknow) is flat. We now conclude that s<strong>in</strong>ce the Gaussian curvature of a sphere is nonzero, a spherecannot be locally isometric to a plane. Thus, there is no way to map the earth “faithfully” (preserv<strong>in</strong>gdistance)—even locally—on a piece of paper. In some sense, the Mercator projection (seeExercise 2.1.9) is the best we can do, for, although it distorts distances, it does preserve angles.Let’s now apply the Codazzi equations to prove a rather strik<strong>in</strong>g result about surfaces withK =0everywhere, called flat surfaces.Proposition 3.3. Suppose M is a flat surface with no planar po<strong>in</strong>ts. Then M is a ruled surfacewhose tangent plane is constant along the rul<strong>in</strong>gs.Proof. S<strong>in</strong>ce M has no planar po<strong>in</strong>ts, we can choose k 1 =0<strong>and</strong> k 2 ≠0everywhere. Thenby Theorem 3.3 of the Appendix, there is a local parametrization of M so that the u-curves arethe first l<strong>in</strong>es of curvature <strong>and</strong> the v-curves are the second l<strong>in</strong>es of curvature. This means first ofall that F = m =0. Now, s<strong>in</strong>ce k 1 =0,for any P ∈ M we have S P (x u )=0, <strong>and</strong> so n u = 0everywhere <strong>and</strong> n is constant along the u-curves.We now want to show that the u-curves are <strong>in</strong> fact l<strong>in</strong>es. <strong>First</strong>, l =II(x u , x u )=−S P (x u )·x u =0<strong>and</strong> n ≠0,s<strong>in</strong>ce k 2 ≠0.From the first of the Codazzi equations we now deduce that<strong>and</strong> so Γ v uu =0. This means that0=l v − m u = lΓ u uv + m ( Γ v uv − Γ u uu)− nΓvuu = −nΓ v uu,x uu =Γ u uux u +Γ v uux v + ln =Γ u uux uis just a multiple of x u .Thus, the tangent vector x u never changes direction as we move along theu-curves, <strong>and</strong> this means that the u-curves must be l<strong>in</strong>es. In conclusion, we have a ruled surfacewhose tangent plane is constant along rul<strong>in</strong>gs. □Remark. Such ruled surfaces are called developable. (See Exercise 10 <strong>and</strong> Exercise 2.1.8.) Theterm<strong>in</strong>ology comes from the fact that they can be rolled out—or “developed”—onto a plane.Next we prove a strik<strong>in</strong>g global result about compact surfaces. (Recall that a subset of R 3 iscompact if it is closed <strong>and</strong> bounded. The salient feature of compact sets is the maximum valuetheorem: A cont<strong>in</strong>uous real-valued function on a compact set achieves its maximum <strong>and</strong> m<strong>in</strong>imumvalues.) We beg<strong>in</strong> with a straightforwardProposition 3.4. Suppose M ⊂ R 3 is a compact surface. Then there is a po<strong>in</strong>t P ∈ M withK(P ) > 0.Proof. Because M is compact, the cont<strong>in</strong>uous function f(x) =‖x‖ achieves its maximum atsome po<strong>in</strong>t of M, <strong>and</strong> so there is a po<strong>in</strong>t P ∈ M farthest from the orig<strong>in</strong> (which may or may not be<strong>in</strong>side M), as <strong>in</strong>dicated <strong>in</strong> Figure 3.2. Let f(P )=R. AsExercise 1.2.7 shows, the curvature of anycurve α ⊂ M at P is at least 1/R, so—if we choose the unit normal n to be <strong>in</strong>ward-po<strong>in</strong>t<strong>in</strong>g—everynormal curvature of M at P is at least 1/R. Itfollows that K(P ) ≥ 1/R 2 > 0. (That is, M is atleast as curved at P as the circumscribed sphere of radius R tangent to M at P .) □


§3. The Codazzi <strong>and</strong> Gauss Equations <strong>and</strong> the Fundamental Theorem of Surface Theory 590MPFigure 3.2The reader is asked <strong>in</strong> Exercise 16 to f<strong>in</strong>d surfaces of revolution of constant curvature. Thereare, <strong>in</strong>terest<strong>in</strong>gly, many nonobvious examples. However, if we restrict ourselves to smooth surfaces,we have the follow<strong>in</strong>g beautifulTheorem 3.5 (Liebmann). If M is a smooth, compact surface of constant Gaussian curvatureK, then K>0 <strong>and</strong> M must be a sphere of radius 1/ √ K.We will need the follow<strong>in</strong>gLemma 3.6 (Hilbert). Suppose P is not an umbilic po<strong>in</strong>t <strong>and</strong> k 1 (P ) >k 2 (P ). Suppose k 1 hasalocal maximum at P <strong>and</strong> k 2 has a local m<strong>in</strong>imum at P . Then K(P ) ≤ 0.Proof. We work <strong>in</strong> a “pr<strong>in</strong>cipal” coord<strong>in</strong>ate parametrization, 4 so that the u-curves are l<strong>in</strong>esof curvature with pr<strong>in</strong>cipal curvature k 1 <strong>and</strong> the v-curves are l<strong>in</strong>es of curvature with pr<strong>in</strong>cipalcurvature k 2 . Then we have k 1 = l/E, k 2 = n/G, <strong>and</strong> F = m =0.Bythe Codazzi equations <strong>and</strong>the equations (‡) onp.55,we have(⋆)l v = 1 2 E v(k 1 + k 2 ) <strong>and</strong> n u = 1 2 G u(k 1 + k 2 ).S<strong>in</strong>ce l = k 1 E,wehavel v =(k 1 ) v E + k 1 E v ; similarly, n u =(k 2 ) u G + k 2 G u . Us<strong>in</strong>g the equations(⋆), we then obta<strong>in</strong>(k 1 ) v = E v2E (k 2 − k 1 ) <strong>and</strong> (k 2 ) u = G u(⋆⋆)2G (k 1 − k 2 );s<strong>in</strong>ce k 1 ≠ k 2 <strong>and</strong> (k 1 ) v =(k 2 ) u =0atP ,we<strong>in</strong>fer that E v = G u =0atP .Differentiat<strong>in</strong>g the equations (⋆⋆), <strong>and</strong> remember<strong>in</strong>g that (k 1 ) u =(k 2 ) v =0atP as well, wehave at P :(k 1 ) vv = E vv2E (k 2 − k 1 ) ≤ 0 (because k 1 has a local maximum at P )(k 2 ) uu = G uu2G (k 1 − k 2 ) ≥ 0 (because k 2 has a local m<strong>in</strong>imum at P ),4 S<strong>in</strong>ce locally there are no umbilic po<strong>in</strong>ts, the existence of such a parametrization is an immediate consequenceof Theorem 3.3 of the Appendix.


60 Chapter 2. <strong>Surfaces</strong>: Local Theory<strong>and</strong> so E vv ≥ 0 <strong>and</strong> G uu ≥ 0atP . Us<strong>in</strong>g the equation (∗) for the Gaussian curvature on p. 57, wesee that−2KEG = E vv + G uu + a(u, v)E v + b(u, v)G ufor some functions a(u, v) <strong>and</strong> b(u, v). So we conclude that K(P ) ≤ 0, as desired.Proof of Theorem 3.5. By Proposition 3.4, there is a po<strong>in</strong>t where M is positively curved,<strong>and</strong> s<strong>in</strong>ce the Gaussian curvature is constant, we must have K>0. If every po<strong>in</strong>t is umbilic, thenby Exercise 2.2.12, we know that M is a sphere. If there is some non-umbilic po<strong>in</strong>t, the largerpr<strong>in</strong>cipal curvature, k 1 ,achieves its maximum value at some po<strong>in</strong>t P because M is compact. Then,s<strong>in</strong>ce K = k 1 k 2 is constant, the function k 2 = K/k 1 must achieve its m<strong>in</strong>imum at P . S<strong>in</strong>ce Pis necessarily a non-umbilic po<strong>in</strong>t (why?), it follows from Lemma 3.6 that K(P ) ≤ 0, which is acontradiction. □Remark. Hopf proved a stronger result, which requires techniques from complex analysis: IfM is a compact surface topologically equivalent to a sphere <strong>and</strong> hav<strong>in</strong>g constant mean curvature,then M must be a sphere.We conclude this section with the analogue of Theorem 3.1 of Chapter 1.Theorem 3.7 (Fundamental Theorem of Surface Theory). Uniqueness: Two parametrized surfacesx, x ∗ : U → R 3 are congruent (i.e., differ by a rigid motion) if <strong>and</strong> only if I = I ∗ <strong>and</strong>II = ±II ∗ . Existence: Moreover, given differentiable functions E, F , G, l, m, <strong>and</strong> n with E>0 <strong>and</strong>EG− F 2 > 0 <strong>and</strong> satisfy<strong>in</strong>g the Codazzi <strong>and</strong> Gauss equations, there exists (locally) a parametrizedsurface x(u, v) with the respective I <strong>and</strong> II.Proof. The existence statement requires some theorems from partial differential equationsbeyond our reach at this stage. The uniqueness statement, however, is much like the proof ofTheorem 3.1 of Chapter 1. (The ma<strong>in</strong> technical difference is that we no longer are lucky enoughto be work<strong>in</strong>g with an orthonormal basis at each po<strong>in</strong>t, as we were with the Frenet frame.)<strong>First</strong>, suppose x ∗ =Ψ◦x for some rigid motion Ψ: R 3 → R 3 (i.e., Ψ(x) =Ax + b for someb ∈ R 3 <strong>and</strong> some 3×3 orthogonal matrix A). S<strong>in</strong>ce a translation doesn’t change partial derivatives,we may assume that b = 0. Now, s<strong>in</strong>ce orthogonal matrices preserve length <strong>and</strong> dot product, wehave E ∗ = ‖x ∗ u‖ 2 = ‖Ax u ‖ 2 = ‖x u ‖ 2 = E, etc., so I = I ∗ . If det A>0, then n ∗ = An, whereas ifdet A0 <strong>and</strong> the negative when det A0 <strong>and</strong> II ∗ = −II ifdet A


§3. The Codazzi <strong>and</strong> Gauss Equations <strong>and</strong> the Fundamental Theorem of Surface Theory 61v ′ j(t) =3∑q ij v i (t) <strong>and</strong> vj ∗′ (t) =i=13∑q ij vi ∗ (t), j =1, 2, 3.(Note that the coefficient functions p i <strong>and</strong> q ij are the same for both the starred <strong>and</strong> unstarredequations.) If α(0) = α ∗ (0) <strong>and</strong> v i (0) = vi ∗(0), i =1, 2, 3, then α(t) =α∗ (t) <strong>and</strong> v i (t) =vi ∗ (t) forall t ∈ [0,b], i =1, 2, 3.Fix a po<strong>in</strong>t u 0 ∈ U. By compos<strong>in</strong>g x ∗ with a rigid motion, we may assume that at u 0 we havex = x ∗ , x u = x ∗ u, x v = x ∗ v, <strong>and</strong> n = n ∗ (why?). Choose an arbitrary u 1 ∈ U, <strong>and</strong> jo<strong>in</strong> u 0 to u 1by a path u(t), t ∈ [0,b], <strong>and</strong> apply the lemma with α = x◦u, v 1 = x u ◦u, v 2 = x v ◦u, v 3 = n◦u,p i = u ′ i , <strong>and</strong> the q ij prescribed by the equations (†) <strong>and</strong> (††). S<strong>in</strong>ce I = I ∗ <strong>and</strong> II = II ∗ , the sameequations hold for α ∗ = x ∗ ◦u, <strong>and</strong> so x(u 1 )=x ∗ (u 1 )asdesired. That is, the two parametrizedsurfaces are identical. □Proof of Lemma 3.8. Introduce the matrix function of t⎡⎤| | |⎢⎥M(t) = ⎣ v 1 (t) v 2 (t) v 3 (t) ⎦ ,| | |<strong>and</strong> analogously for M ∗ (t). Then the displayed equations <strong>in</strong> the statement of the Lemma can bewritten asi=1M ′ (t) =M(t)Q(t) <strong>and</strong> M ∗′ (t) =M ∗ (t)Q(t).On the other h<strong>and</strong>, we have M(t) T M(t) =G(t). S<strong>in</strong>ce the v i (t) form a basis for R 3 for eacht, weknow the matrix G is <strong>in</strong>vertible. Now, differentiat<strong>in</strong>g the equation G(t)G −1 (t) =I yields(G −1 ) ′ (t) =−G −1 (t)G ′ (t)G −1 (t), <strong>and</strong> differentiat<strong>in</strong>g the equation G(t) =M(t) T M(t) yields G ′ (t) =M ′ (t) T M(t)+M(t) T M ′ (t) =Q(t) T G(t)+G(t)Q(t). Now consider(M ∗ G −1 M T ) ′ (t) =M ∗′ (t)G(t) −1 M(t) T + M ∗ (t)(G −1 ) ′ (t)M(t) T + M ∗ (t)G(t) −1 M ′ (t) T= M ∗ (t)Q(t)G(t) −1 M(t) T + M ∗ (t) ( −G(t) −1 G ′ (t)G(t) −1) M(t) T+ M ∗ (t)G(t) −1 Q(t) T M(t) T= M ∗ (t)Q(t)G(t) −1 M(t) T − M ∗ (t)G(t) −1 Q(t) T M(t) T − M ∗ (t)Q(t)G(t) −1 M(t) T+ M ∗ (t)G(t) −1 Q(t) T M(t) T = O.S<strong>in</strong>ce M(0) = M ∗ (0), we have M ∗ (0)G(0) −1 M(0) T = M(0)M(0) −1 M(0) T−1 M(0) T = I, <strong>and</strong> soM ∗ (t)G(t) −1 M(t) T = I for all t ∈ [0,b]. It follows that M ∗ (t) =M(t) for all t ∈ [0,b], <strong>and</strong> soα ∗′ (t) − α ′ (t) =0 for all t as well. S<strong>in</strong>ce α ∗ (0) = α(0), it follows that α ∗ (t) =α(t) for all t ∈ [0,b],as we wished to establish. □EXERCISES 2.31. Calculate the Christoffel symbols for a cone, x(u, v) =(u cos v, u s<strong>in</strong> v, u), both directly <strong>and</strong> byus<strong>in</strong>g the formulas (‡).


62 Chapter 2. <strong>Surfaces</strong>: Local Theory2. Use the first Gauss equation to derive the formula (∗) given on p. 57 for Gaussian curvature.3. Check the Gaussian curvature of the sphere us<strong>in</strong>g the formula (∗) onp.57.4. Check that for a parametrized surface with E = G = λ(u, v) <strong>and</strong> F =0,the Gaussian curvatureis given by K = − 12λ ∇2 (ln λ). (Here ∇ 2 f = ∂2 f∂u 2 + ∂2 fis the Laplacian of f.)∂v2 5. Calculate the Christoffel symbols for the follow<strong>in</strong>g parametrized surfaces. Then check <strong>in</strong> eachcase that the Codazzi equations <strong>and</strong> the first Gauss equation hold.a. the plane, parametrized by polar coord<strong>in</strong>ates: x(u, v) =(u cos v, u s<strong>in</strong> v, 0)b. a helicoid: x(u, v) =(u cos v, u s<strong>in</strong> v, v)♯ c. a cone: x(u, v) =(u cos v, u s<strong>in</strong> v, cu), c ≠0♯ *d. a surface of revolution: x(u, v) = ( f(u) cos v, f(u) s<strong>in</strong> v, g(u) ) , with f ′ (u) 2 + g ′ (u) 2 =16. Prove there is no compact m<strong>in</strong>imal surface M ⊂ R 3 .7. Decide whether there is a parametrized surface x(u, v) witha. E = G =1,F =0,l =1=−n, m =0b. E = G =1,F =0,l = e u = n, m =0c. E =1,F =0,G = cos 2 u, l = cos 2 u, m =0,n =18. a. Modify the proof of Theorem 3.5 to prove that a smooth, compact surface with K>0 <strong>and</strong>constant mean curvature is a sphere.b. Give an example to show that the result of Lemma 3.6 fails if we assume k 1 has a localm<strong>in</strong>imum <strong>and</strong> k 2 has a local maximum at P .9. Give examples of (locally) non-congruent parametrized surfaces x <strong>and</strong> x ∗ witha. I=I ∗b. II = II ∗ (H<strong>in</strong>t: Try reparametriz<strong>in</strong>g some of our simplest surfaces.)10. Let x(u, v) =α(u) +vβ(u) beaparametrization of a ruled surface. Prove that the tangentplane is constant along rul<strong>in</strong>gs (i.e., the surface is flat) if <strong>and</strong> only if α ′ (u), β(u), <strong>and</strong> β ′ (u) arel<strong>in</strong>early dependent for every u. (H<strong>in</strong>t: When is S(x v )=0? Alternatively, consider x u × x v <strong>and</strong>apply Exercise A.2.1.)11. Prove that α is a l<strong>in</strong>e of curvature <strong>in</strong> M if <strong>and</strong> only if the ruled surface formed by the surfacenormals along α is flat. (H<strong>in</strong>t: See Exercise 10.)12. Show that the Gaussian curvature of the parametrized surfacesx(u, v) =(u cos v, u s<strong>in</strong> v, v)y(u, v) =(u cos v, u s<strong>in</strong> v, ln u)is the same for each (u, v), <strong>and</strong> yet the first fundamental forms I x <strong>and</strong> I y do not agree. (Thus,the converse of Corollary 3.2 is false.)13. Suppose that the surface M is doubly ruled by orthogonal l<strong>in</strong>es (i.e., through each po<strong>in</strong>t of Mthere pass two orthogonal l<strong>in</strong>es).


§4. Covariant Differentiation, Parallel Translation, <strong>and</strong> Geodesics 63a. Us<strong>in</strong>g the Gauss equations, prove that K =0.b. Now deduce that M must be a plane.(H<strong>in</strong>t: As usual, assume the families of l<strong>in</strong>es are u- <strong>and</strong> v-curves.)14. Suppose M is a surface with no umbilic po<strong>in</strong>ts <strong>and</strong> one constant pr<strong>in</strong>cipal curvature k 1 ≠0.Prove that M is (a subset of) a tube of radius r =1/|k 1 | about a curve. That is, there is acurve α so that M is (a subset of) the union of circles of radius r <strong>in</strong> each normal plane, centeredalong the curve. (H<strong>in</strong>ts: As usual, work with a parametrization where the u-curves are l<strong>in</strong>esof curvature with pr<strong>in</strong>cipal curvature k 1 <strong>and</strong> the v-curves are l<strong>in</strong>es of curvature with pr<strong>in</strong>cipalcurvature k 2 . Use the Codazzi equations to show that the u-curves have curvature |k 1 | <strong>and</strong> areplanar. Then def<strong>in</strong>e α appropriately <strong>and</strong> check that it is a regular curve.)15. Consider the parametrized surfacesx(u, v) =(− cosh u s<strong>in</strong> v, cosh u cos v, u)y(u, v) =(u cos v, u s<strong>in</strong> v, v)(a helicoid).(a catenoid)a. Compute the first <strong>and</strong> second fundamental forms of both surfaces, <strong>and</strong> check that bothsurfaces are m<strong>in</strong>imal.b. F<strong>in</strong>d the asymptotic curves on both surfaces.c. Show that we can locally reparametrize the helicoid <strong>in</strong> such a way as to make the firstfundamental forms of the two surfaces agree; this means that the two surfaces are locallyisometric. (H<strong>in</strong>t: See p. 39. Replace u with s<strong>in</strong>h u <strong>in</strong> the parametrization of the helicoid.Why is this legitimate?)d. Why are they not globally isometric?e. (for the student who’s seen a bit of complex variables) As a h<strong>in</strong>t to what’s go<strong>in</strong>g on here,let z = u + iv <strong>and</strong> Z = x + iy, <strong>and</strong> check that, cont<strong>in</strong>u<strong>in</strong>g to use the substitution frompart c, Z = (s<strong>in</strong> iz, cos iz, z). Underst<strong>and</strong> now how one can obta<strong>in</strong> a one-parameter familyof isometric surfaces <strong>in</strong>terpolat<strong>in</strong>g between the helicoid <strong>and</strong> the catenoid.16. F<strong>in</strong>d all the surfaces of revolution of constant curvaturea. K =0b. K =1c. K = −1(H<strong>in</strong>t: There are more than you might suspect. But your answers will <strong>in</strong>volve <strong>in</strong>tegrals youcannot express <strong>in</strong> terms of elementary functions.)4. Covariant Differentiation, Parallel Translation, <strong>and</strong> GeodesicsNow we turn to the “<strong>in</strong>tr<strong>in</strong>sic” geometry of a surface, i.e., the geometry that can be observed byan <strong>in</strong>habitant (for example, a flat ant) of the surface, who can only perceive what happens along (or,say, tangential to) the surface. Anyone who has studied Euclidean geometry knows how importantthe notion of parallelism is (<strong>and</strong> classical non-Euclidean geometry arises when one removes Euclid’s


64 Chapter 2. <strong>Surfaces</strong>: Local Theoryparallel postulate, which stipulates that given any l<strong>in</strong>e L <strong>in</strong> the plane <strong>and</strong> any po<strong>in</strong>t P not ly<strong>in</strong>g onL, there is a unique l<strong>in</strong>e through P parallel to L). It seems quite <strong>in</strong>tuitive to say that, work<strong>in</strong>g just<strong>in</strong> R 3 ,two vectors V (thought of as be<strong>in</strong>g “tangent at P ”) <strong>and</strong> W (thought of as be<strong>in</strong>g “tangent atQ”) are parallel provided that we obta<strong>in</strong> W when we move V “parallel to itself” from P to Q; <strong>in</strong>other words, if W = V. But what would an <strong>in</strong>habitant of the sphere say? How should he compareVWPQAre V <strong>and</strong> W parallel?Figure 4.1a tangent vector at one po<strong>in</strong>t of the sphere to a tangent vector at another <strong>and</strong> determ<strong>in</strong>e if they’re“parallel”? (See Figure 4.1.) Perhaps a better question is this: Given a curve α on the surface <strong>and</strong>avector field X def<strong>in</strong>ed along α, should we say X is parallel if it has zero derivative along α?We already know how an <strong>in</strong>habitant differentiates a scalar function f : M → R, byconsider<strong>in</strong>gthe directional derivative D V f for any tangent vector V ∈ T P M.Wenow beg<strong>in</strong> with aDef<strong>in</strong>ition. We say a function X: M → R 3 is a vector field on M if(1) X(P ) ∈ T P M for every P ∈ M, <strong>and</strong>(2) for any parametrization x: U → M, the function X◦x: U → R 3 is (cont<strong>in</strong>uously) differentiable.Now, we can differentiate a vector field X on M <strong>in</strong> the customary fashion: If V ∈ T P M,wechoose a curve α with α(0) = P <strong>and</strong> α ′ (0) = V <strong>and</strong> set D V X =(X◦α) ′ (0). (As usual, the cha<strong>in</strong>rule tells us this is well-def<strong>in</strong>ed.) But the <strong>in</strong>habitant of the surface can only see that portion of thisvector ly<strong>in</strong>g <strong>in</strong> the tangent plane. This br<strong>in</strong>gs us to theDef<strong>in</strong>ition. Given a vector field X <strong>and</strong> V ∈ T P M,wedef<strong>in</strong>e the covariant derivative∇ V X =(D V X) ‖ = the projection of D V X onto T P M = D V X − (D V X · n)n.Given a curve α <strong>in</strong> M, wesay the vector field X is covariant constant or parallel along α if∇ α ′ (t)X = 0 for all t. (This means that D α ′ (t)X =(X◦α) ′ (t) isamultiple of the normal vectorn(α(t)).)Example 1. Let M be asphere <strong>and</strong> let α be agreat circle <strong>in</strong> M. The derivative of the unittangent vector of α po<strong>in</strong>ts towards the center of the circle, which is <strong>in</strong> this case the center of thesphere, <strong>and</strong> thus is completely normal to the sphere. Therefore, the unit tangent vector field ofα is parallel along α. Observe that the constant vector field (0, 0, 1) is parallel along the equatorz =0of a sphere centered at the orig<strong>in</strong>. Is this true of any other constant vector field? ▽


§4. Covariant Differentiation, Parallel Translation, <strong>and</strong> Geodesics 65Example 2. A fundamental example requires that we revisit the Christoffel symbols. Given aparametrized surface x: U → M, wehave∇ xu x u =(x uu ) ‖ =Γ u uux u +Γ v uux v∇ xv x u =(x uv ) ‖ =Γ u uvx u +Γ v uvx v = ∇ xu x v ,∇ xv x v =(x vv ) ‖ =Γ uvvx u +Γ vvvx v .▽<strong>and</strong>The first result we prove is the follow<strong>in</strong>gProposition 4.1. Given a curve α: [0, 1] → M with α(0) = P <strong>and</strong> X 0 ∈ T P M, there is aunique parallel vector field X def<strong>in</strong>ed along α with X(P )=X 0 .Proof. Assum<strong>in</strong>g α lies <strong>in</strong> a parametrized portion x: U → M, set α(t) =x(u(t),v(t)) <strong>and</strong>write X(α(t)) = a(t)x u (u(t),v(t)) + b(t)x v (u(t),v(t)). Then α ′ (t) =u ′ (t)x u + v ′ (t)x v (where thethe cumbersome argument (u(t),v(t)) is understood). So, by the product rule <strong>and</strong> cha<strong>in</strong> rule, wehave∇ α ′ (t)X = ( (X◦α) ′ (t) ) (‖ d (= a(t)xu (u(t),v(t)) + b(t)x v (u(t),v(t)) )) ‖dt( d ‖ ( ) d ‖= a ′ (t)x u + b ′ (t)x v + a(t) u(u(t),v(t)))dt x + b(t)dt x v(u(t),v(t))= a ′ (t)x u + b ′ (t)x v + a(t) ( u ′ (t)x uu + v ′ ) ‖ ((t)x uv + b(t) u ′ (t)x vu + v ′ ) ‖(t)x vv= a ′ (t)x u + b ′ (t)x v + a(t) ( u ′ (t)(Γuux u u +Γuux v v )+v ′ (t)(Γuvx u u +Γuvx v v ) )+ b(t) ( u ′ (t)(Γvux u u +Γvux v v )+v ′ (t)(Γvvx u u +Γvvx v v ) )= ( a ′ (t)+a(t)(Γuuu u ′ (t)+Γuvv u ′ (t)) + b(t)(Γvuu u ′ (t)+Γvvv u ′ (t)) ) x u+ ( b ′ (t)+a(t)(Γuuu v ′ (t)+Γuvv v ′ (t)) + b(t)(Γvuu v ′ (t)+Γvvv v ′ (t)) ) x v .Thus, to say X is parallel along the curve α is to say that a(t) <strong>and</strong> b(t) are solutions of the l<strong>in</strong>earsystem of first order differential equationsa ′ (t)+a(t)(Γuuu u ′ (t)+Γuvv u ′ (t)) + b(t)(Γvuu u ′ (t)+Γvvv u ′ (t)) = 0(♣)b ′ (t)+a(t)(Γuuu v ′ (t)+Γuvv v ′ (t)) + b(t)(Γvuu v ′ (t)+Γvvv v ′ (t)) =0.By Theorem 3.2 of the Appendix, this system has a unique solution on [0, 1] once we specify a(0)<strong>and</strong> b(0), <strong>and</strong> hence we obta<strong>in</strong> a unique parallel vector field X with X(P )=X 0 . □Def<strong>in</strong>ition. If Q = α(1), we refer to X(Q) astheparallel translate of X 0 along α, orthe resultof parallel translation along α.Example 3. Fix a latitude circle u = u 0 (u 0 ≠0,π)onthe unit sphere <strong>and</strong> let’s calculate theeffect of parallel-translat<strong>in</strong>g the vector X 0 = x v start<strong>in</strong>g at the po<strong>in</strong>t P given by u = u 0 , v =0,oncearound the circle, counterclockwise. We parametrize the curve by u(t) =u 0 , v(t) =t, 0≤ t ≤ 2π.Us<strong>in</strong>g our computation of the Christoffel symbols of the sphere <strong>in</strong> Example 1 or 2 of Section 3, weobta<strong>in</strong> from (♣) the differential equationsa ′ (t) =s<strong>in</strong> u 0 cos u 0 b(t), a(0) = 0


66 Chapter 2. <strong>Surfaces</strong>: Local Theoryb ′ (t) =− cot u 0 a(t), b(0) =1.We solve this system by differentiat<strong>in</strong>g the second equation aga<strong>in</strong> <strong>and</strong> substitut<strong>in</strong>g the first:Thus, the solution isb ′′ (t) =− cot u 0 a ′ (t) =− cos 2 u 0 b(t), b(0) = 1.a(t) =s<strong>in</strong> u 0 s<strong>in</strong> ( (cos u 0 )t ) , b(t) =cos ( (cos u 0 )t ) .Note that ‖X(α(t))‖ 2 = Ea(t) 2 +2Fa(t)b(t)+Gb(t) 2 = s<strong>in</strong> 2 u 0 for all t. That is, the orig<strong>in</strong>al vectorX 0 rotates as we parallel translate it around the latitude circle, <strong>and</strong> its length is preserved. Aswe see <strong>in</strong> Figure 4.2, the vector rotates clockwise as we proceed around the latitude circle (<strong>in</strong> theP2π cos u 0u 0X 0Figure 4.2upper hemisphere). But this makes sense: If we just take the covariant derivative of the tangentvector to the circle, it po<strong>in</strong>ts upwards (cf. Figure 3.1), so the vector field must rotate clockwise tocounteract that effect <strong>in</strong> order to rema<strong>in</strong> parallel. ▽Example 4 (Foucault pendulum). Foucault observed <strong>in</strong> 1851 that the sw<strong>in</strong>g plane of a pendulumlocated on the latitude circle u = u 0 precesses with a period of T =24/ cos u 0 hours. Wecan use the result of Example 3 to expla<strong>in</strong> this. We imag<strong>in</strong>e the earth as fixed <strong>and</strong> “transport” thesw<strong>in</strong>g<strong>in</strong>g pendulum once around the circle <strong>in</strong> 24 hours. If we make the pendulum very long <strong>and</strong>the sw<strong>in</strong>g rather short, the motion will be “essentially” tangential to the surface of the earth. If wemove slowly around the circle, the forces will be “essentially” normal to the sphere: In particular,lett<strong>in</strong>g R denote the radius of the earth, the tangential component of the centripetal accelerationis (cf. Figure 3.1)( ) 2π 2(R s<strong>in</strong> u 0 ) cos u 0 ≤ 2π2 R24 24 2 ≈ 0.0558 ft/sec 2 ≈ 0.17%g.Thus, the “sw<strong>in</strong>g vector field” is, for all practical purposes, parallel along the curve. Therefore, it2πturns through an angle of 2π cos u 0 <strong>in</strong> one trip around the circle, so it takes(2π cos u 0 )/24 = 24cos u 0hours to return to its orig<strong>in</strong>al sw<strong>in</strong>g plane. ▽


§4. Covariant Differentiation, Parallel Translation, <strong>and</strong> Geodesics 67Our experience <strong>in</strong> Example 3 suggests the follow<strong>in</strong>gProposition 4.2. Parallel translation preserves lengths <strong>and</strong> angles. That is, if X <strong>and</strong> Y areparallel vector fields along a curve α from P to Q, then ‖X(P )‖ = ‖X(Q)‖ <strong>and</strong> the angle betweenX(P ) <strong>and</strong> Y(P ) equals the angle between X(Q) <strong>and</strong> Y(Q).Proof. Consider f(t) =X(α(t)) · Y(α(t)). Thenf ′ (t) =(X◦α) ′ (t) · (Y◦α)(t)+(X◦α)(t) · (Y◦α) ′ (t)= D α ′ (t)X · Y + X · D α ′ (t)Y (1)= ∇ α ′ (t)X · Y + X ·∇ α ′ (t)Y (2)=0.Note that equality (1) holds because X <strong>and</strong> Y are tangent to M <strong>and</strong> hence their dot product withany vector normal to the surface is 0. Equality (2) holds because X <strong>and</strong> Y are assumed parallel alongα. Itfollows that the dot product X·Y rema<strong>in</strong>s constant along α. Tak<strong>in</strong>g Y = X, we<strong>in</strong>fer that ‖X‖(<strong>and</strong> similarly ‖Y‖) isconstant. Know<strong>in</strong>g that, us<strong>in</strong>g the famous formula cos θ = X · Y/‖X‖‖Y‖for the angle θ between X <strong>and</strong> Y, we<strong>in</strong>fer that the angle rema<strong>in</strong>s constant. □Now we change gears somewhat. We saw <strong>in</strong> Exercise 1.1.8 that the shortest path jo<strong>in</strong><strong>in</strong>g twopo<strong>in</strong>ts <strong>in</strong> R 3 is a l<strong>in</strong>e segment. One characterization of the l<strong>in</strong>e segment is that it never changesdirection, so that its unit tangent vector is parallel (so no distance is wasted by turn<strong>in</strong>g). It seemsplausible that the mythical <strong>in</strong>habitant of our surface M might try to travel from one po<strong>in</strong>t toanother <strong>in</strong> M, stay<strong>in</strong>g <strong>in</strong> M, bysimilarly not turn<strong>in</strong>g; that is, so that his unit tangent vectorfield is parallel along his path. Physically, this means that if he travels at constant speed, anyacceleration should be normal to the surface. This leads us to the follow<strong>in</strong>gDef<strong>in</strong>ition. We say a parametrized curve α <strong>in</strong> a surface M is a geodesic if its tangent vectoris parallel along the curve, i.e., if ∇ α ′α ′ =0.Recall that s<strong>in</strong>ce parallel translation preserves lengths, α must have constant speed, although itmay not be arclength-parametrized. In general, we refer to an unparametrized curve as a geodesicif its arclength-reparametrization is <strong>in</strong> fact a geodesic.In general, given any arclength-parametrized curve α ly<strong>in</strong>g on M, wedef<strong>in</strong>ed its normal curvatureat the end of Section 2. Instead of us<strong>in</strong>g the Frenet frame, it is natural to consider the Darbouxframe for α, which takes <strong>in</strong>to account the fact that α lies on the surface M. (Both are illustrated<strong>in</strong> Figure 4.3.) We take the right-h<strong>and</strong>ed orthonormal basis {T, n × T, n}; note that the first twoBn×TnNTThe Frenet <strong>and</strong> Darboux framesFigure 4.3vectors give a basis for T P M.Wecan decompose the curvature vectorκN = ( ) ( )κN · (n × T) (n × T)+ κN} {{ }} {{ · n } n.κ gκ n


68 Chapter 2. <strong>Surfaces</strong>: Local TheoryAs we saw before, κ n gives the normal component of the curvature vector; κ g gives the tangentialcomponent of the curvature vector <strong>and</strong> is called the geodesic curvature. This term<strong>in</strong>ology arisesfrom the fact that α is a geodesic if <strong>and</strong> only if its geodesic curvature vanishes.Example 5. We saw <strong>in</strong> Example 1 that every great circle on a sphere is a geodesic. Are thereothers? Let α be a geodesic on a sphere centered at the orig<strong>in</strong>. S<strong>in</strong>ce κ g =0,the accelerationvector α ′′ (s) must be a multiple of α(s) for every s, <strong>and</strong> so α ′′ × α = 0. Therefore α ′ × α = A isa constant vector, so α lies <strong>in</strong> the plane pass<strong>in</strong>g through the orig<strong>in</strong> with normal vector A. That is,α is a great circle. ▽Us<strong>in</strong>g the equations (♣), let’s now give the equations for the curve α(t) =x(u(t),v(t)) to bea geodesic. S<strong>in</strong>ce X = α ′ (t) =u ′ (t)x u + v ′ (t)x v ,wehave a(t) =u ′ (t) <strong>and</strong> b(t) =v ′ (t), <strong>and</strong> theresult<strong>in</strong>g equations are(♣♣)u ′′ (t)+Γ u uuu ′ (t) 2 +2Γ u uvu ′ (t)v ′ (t)+Γ uvvv ′ (t) 2 =0v ′′ (t)+Γ v uuu ′ (t) 2 +2Γ v uvu ′ (t)v ′ (t)+Γ vvvv ′ (t) 2 =0.The follow<strong>in</strong>g result is a consequence of basic results on differential equations (see Theorem 3.1of the Appendix).Proposition 4.3. Given a po<strong>in</strong>t P ∈ M <strong>and</strong> V ∈ T P M, V ≠ 0, there exist ε>0 <strong>and</strong> a uniquegeodesic α: (−ε, ε) → M with α(0) = P <strong>and</strong> α ′ (0) = V.Example 6. We now use the equations (♣♣) tosolve for geodesics analytically <strong>in</strong> a few examples.(a) Let x(u, v) =(u, v) bethe obvious parametrization of the plane. Then all the Christoffelsymbols vanish <strong>and</strong> the geodesics are the solutions ofu ′′ (t) =v ′′ (t) =0,so we get the l<strong>in</strong>es α(t) =(u(t),v(t)) = (a 1 t + b 1 ,a 2 t + b 2 ), as expected. Note that α does<strong>in</strong> fact have constant speed.(b) Us<strong>in</strong>g the st<strong>and</strong>ard spherical coord<strong>in</strong>ate parametrization of the sphere, we obta<strong>in</strong> (see Example1 or 2 of Section 3) the equations(∗)u ′′ (t) − s<strong>in</strong> u(t) cos u(t)v ′ (t) 2 =0=v ′′ (t)+2cotu(t)u ′ (t)v ′ (t).Well, one obvious set of solutions is to take u(t) =t, v(t) =v 0 (<strong>and</strong> these, <strong>in</strong>deed, givethe great circles through the north pole). More generally, let’s now impose the additionalcondition that the curve be arclength-parametrized:(∗∗)u ′ (t) 2 + s<strong>in</strong> 2 u(t)v ′ (t) 2 =1.Integrat<strong>in</strong>g the equation <strong>in</strong> (∗) weobta<strong>in</strong> ln v ′ (t) =−2lns<strong>in</strong>u(t)+const, sov ′ c(t) =s<strong>in</strong> 2 u(t)


§4. Covariant Differentiation, Parallel Translation, <strong>and</strong> Geodesics 69for some constant c. Now, us<strong>in</strong>g (∗∗), <strong>and</strong> switch<strong>in</strong>g to Leibniz notation for obvious reasons,we obta<strong>in</strong>√dudt = ± 1 −dtc 2s<strong>in</strong> 2 u(t) ,dvdvdu = dt c csc 2 u= ± √du 1 − c 2 csc 2 u ;<strong>and</strong> sothus, separat<strong>in</strong>g variables givesdv = ±c csc2 udu√1 − c 2 csc 2 u = ± c csc 2 udu√(1 − c 2 ) − c 2 cot 2 u .Now we make the substitution c cot u = √ 1 − c 2 s<strong>in</strong> w; then we havec csc 2 ududv = ± √(1 − c 2 ) − c 2 cot 2 u = ∓dw,<strong>and</strong> so, at long last, we have w = ±v + a for some constant a. Thus,c cot u = √ 1 − c 2 s<strong>in</strong> w = √ 1 − c 2 s<strong>in</strong>(±v + a) = √ 1 − c 2 (s<strong>in</strong> a cos v ± cos a s<strong>in</strong> v),<strong>and</strong> so, f<strong>in</strong>ally, we have the equationc cos u + √ 1 − c 2 s<strong>in</strong> u(A cos v + B s<strong>in</strong> v) =0,which we should recognize as the equation of a great circle! (This curve lies on the plane√1 − c 2 (Ax + By)+cz = 0.) ▽We can now give a beautiful geometric description of the geodesics on a surface of revolution.Proposition 4.4 (Clairaut’s relation). The geodesics on a surface of revolution satisfy theequation(♦)r cos φ = const,where r is the distance from the axis of revolution <strong>and</strong> φ is the angle between the geodesic <strong>and</strong> theparallel. Conversely, any curve satisfy<strong>in</strong>g (♦) that is not a parallel is a geodesic.Proof. For the surface of revolution parametrized as <strong>in</strong> Example 7 of Section 2, we have E =1,F =0,G = f(u) 2 ,Γuv v =Γvu v = f ′ (u)/f(u), Γvv u = −f(u)f ′ (u), <strong>and</strong> all other Christoffel symbolsare 0 (see Exercise 2.3.5d.). Then the system (♣♣) ofdifferential equations becomes(† 1 )(† 2 )u ′′ − ff ′ (v ′ ) 2 =0v ′′ + 2f ′f u′ v ′ =0.Rewrit<strong>in</strong>g the equation († 2 ) <strong>and</strong> <strong>in</strong>tegrat<strong>in</strong>g, we obta<strong>in</strong>v ′′ (t)′v ′ (t) = −2f (u(t))u ′ (t)f(u(t))ln v ′ (t) =−2lnf(u(t)) + constv ′ c(t) =f(u(t)) 2 ,


70 Chapter 2. <strong>Surfaces</strong>: Local Theoryso along a geodesic the quantity f(u) 2 v ′ = Gv ′ is constant. We recognize this as the dot productof the tangent vector of our geodesic with the vector x v , <strong>and</strong> so we <strong>in</strong>fer that ‖x v ‖ cos φ = r cos φis constant. (Recall that the tangent vector of the geodesic has constant length.)To this po<strong>in</strong>t we have seen that the equation († 2 )isequivalent to the condition r cos φ = const,provided we assume ‖α ′ ‖ 2 = u ′2 + Gv ′2 is constant as well. But ifwe differentiate <strong>and</strong> obta<strong>in</strong>u ′ (t) 2 + Gv ′ (t) 2 = u ′ (t) 2 + f(u(t)) 2 v ′ (t) 2 = const,u ′ (t)u ′′ (t)+f(u(t)) 2 v ′ (t)v ′′ (t)+f(u(t))f ′ (u(t))u ′ (t)v ′ (t) 2 =0;substitut<strong>in</strong>g for v ′′ (t) us<strong>in</strong>g († 2 ), we f<strong>in</strong>du ′ (t) ( u ′′ (t) − f(u(t))f ′ (u(t))v ′ (t) 2) =0.In other words, provided u ′ (t) ≠0,aconstant-speed curve satisfy<strong>in</strong>g († 2 ) satisfies († 1 )aswell. (SeeExercise 5 for the case of the parallels.) □Remark. We can give a simple physical <strong>in</strong>terpretation of Clairaut’s relation. Imag<strong>in</strong>e a particleconstra<strong>in</strong>ed to move along a surface. If no external forces are act<strong>in</strong>g, then angular momentum isconserved as the particle moves along a geodesic. In the case of our surface of revolution, the verticalcomponent of the angular momentum L = α × α ′ is—surprise, surprise!—f 2 v ′ , which we’ve shownis constant. Perhaps some forces normal to the surface are required to keep the particle on thesurface; then the result<strong>in</strong>g torques have no vertical component, <strong>in</strong>asmuch as (α × n) · (0, 0, 1) = 0.Return<strong>in</strong>g to our orig<strong>in</strong>al motivation for geodesics, we now prove the follow<strong>in</strong>gTheorem 4.5. Geodesics are locally distance-m<strong>in</strong>imiz<strong>in</strong>g.Proof. Choose P ∈ M arbitrary <strong>and</strong> a geodesic γ through P . Start by draw<strong>in</strong>g a curve C 0through P orthogonal to γ. Wenow choose a parametrization x(u, v) sothat x(0, 0) = P , theQv-curvesγPu-curvesC 0Figure 4.4u-curves are geodesics orthogonal to C 0 , <strong>and</strong> the v-curves are the orthogonal trajectories of theu-curves. (It follows from Theorem 3.3 of the Appendix that we can do this on some neighborhoodof P .) We wish to show that for any po<strong>in</strong>t Q = x(u 0 , 0) on γ, any path from P to Q is at least aslong as the length of the geodesic segment.In this parametrization we have F =0<strong>and</strong> E = E(u) (see Exercise 10). Now, if α(t) =x(u(t),v(t)), a ≤ t ≤ b, isany path from P = x(0, 0) to Q = x(u 0 , 0), we have∫ b √∫ b √length(α) = E(u(t))u ′ (t) 2 + G(u(t),v(t))v ′ (t) 2 dt ≥ E(u(t))|u ′ (t)|dtaa


§4. Covariant Differentiation, Parallel Translation, <strong>and</strong> Geodesics 71≥∫ u00√E(u)du,which is the length of the geodesic arc γ from P to Q.□(Cf. Exercise 1.3.1.)Example 7. Why is Theorem 4.5 a local statement? Well, consider a great circle on a sphere,as shown <strong>in</strong> Figure 4.5. If we go more than halfway around, we obviously have not taken theshortest path. ▽shortestPQlongerFigure 4.5Remark. It turns out that any surface can be endowed with a metric (or distance measure)by def<strong>in</strong><strong>in</strong>g the distance between any two po<strong>in</strong>ts to be the <strong>in</strong>fimum (usually, the m<strong>in</strong>imum) of thelengths of all piecewise-C 1 paths jo<strong>in</strong><strong>in</strong>g them. (Although the distance measure is different fromthe Euclidean distance as the surface sits <strong>in</strong> R 3 , the topology—notion of “neighborhood”—<strong>in</strong>ducedby this metric structure is the <strong>in</strong>duced topology that the surface <strong>in</strong>herits as a subspace of R 3 .) It isa consequence of the Hopf-R<strong>in</strong>ow Theorem (see M. doCarmo, Differential Geometry of <strong>Curves</strong> <strong>and</strong><strong>Surfaces</strong>, Prentice Hall, 1976, p. 333, or M. Spivak, A Comprehensive Introduction to DifferentialGeometry, third edition, volume 1, Publish or Perish, Inc., 1999, p. 342) that <strong>in</strong> a surface <strong>in</strong> whichevery parametrized geodesic is def<strong>in</strong>ed for all time (a “complete” surface), every two po<strong>in</strong>ts are <strong>in</strong>fact jo<strong>in</strong>ed by a geodesic of least length. The proof of this result is quite tantaliz<strong>in</strong>g: To f<strong>in</strong>d theshortest path from P to Q, one walks around the “geodesic circle” of po<strong>in</strong>ts a small distance fromP <strong>and</strong> f<strong>in</strong>ds the po<strong>in</strong>t R on it closest to Q; one then proves that the unique geodesic emanat<strong>in</strong>gfrom P that passes through R must eventually pass through Q, <strong>and</strong> there can be no shorter path.We referred earlier to two surfaces M <strong>and</strong> M ∗ as be<strong>in</strong>g globally isometric (e.g., <strong>in</strong> Example 6<strong>in</strong> Section 1). We can now give the official def<strong>in</strong>ition: There should be a function f : M → M ∗ thatestablishes a one-to-one correspondence <strong>and</strong> preserves distance—for any P, Q ∈ M, the distancebetween P <strong>and</strong> Q <strong>in</strong> M should be equal to the distance between f(P ) <strong>and</strong> f(Q) <strong>in</strong>M ∗ .EXERCISES 2.41. Determ<strong>in</strong>e the result of parallel translat<strong>in</strong>g the vector (0, 0, 1) once around the circle x 2 + y 2 =a 2 , z =0,onthe right circular cyl<strong>in</strong>der x 2 + y 2 = a 2 .


72 Chapter 2. <strong>Surfaces</strong>: Local Theory2. Prove that κ 2 = κ 2 g + κ 2 n.3. Suppose α is a non-arclength-parametrized curve. Us<strong>in</strong>g the formula (∗∗) onp.14, prove thatthe velocity vector of α is parallel along α if <strong>and</strong> only if κ g =0<strong>and</strong> υ ′ =0.*4. F<strong>in</strong>d the geodesic curvature κ g of a latitude circle u = u 0 on the unit spherea. directlyb. by apply<strong>in</strong>g the result of Exercise 25. Check that the parallel u = u 0 is a geodesic on the surface of revolution parametrized as <strong>in</strong>Proposition 4.4 if <strong>and</strong> only if f ′ (u 0 )=0. Give a geometric <strong>in</strong>terpretation of <strong>and</strong> explanationfor this result.6. Use the equations (♣) todeterm<strong>in</strong>e through what angle a vector turns when it is paralleltranslatedonce around the circle u = u 0 on the cone x(u, v) =(u cos v, u s<strong>in</strong> v, cu), c ≠0. (SeeExercise 2.3.5c.)7. a. Prove that if the surfaces M <strong>and</strong> M ∗ are tangent along the curve C, parallel translationalong C is the same <strong>in</strong> both surfaces.b. Use the result of part a to determ<strong>in</strong>e the effect of parallel translation around the latitudecircle u = u 0 on the unit sphere. Figure 4.6 <strong>and</strong> basic Euclidean geometry may be of help.(Cf. Example 3.)Figure 4.6*8. What curves ly<strong>in</strong>g on a sphere have constant geodesic curvature?9. Use the equations (♣♣) tof<strong>in</strong>d the geodesics on the plane parametrized by polar coord<strong>in</strong>ates.(H<strong>in</strong>t: Exam<strong>in</strong>e Example 6(b).)♯ 10. a. Suppose F =0<strong>and</strong> the u-curves are geodesics. Use the equations (♣♣) toprove that Eis a function of u only.b. Suppose F =0<strong>and</strong> the u- <strong>and</strong> v-curves are geodesics. Prove that the surface is flat.11. a. Prove that an arclength-parametrized curve α on a surface M with κ ≠0is a geodesic if<strong>and</strong> only if n = ±N.b. Let α be a space curve, <strong>and</strong> let M be the ruled surface generated by its b<strong>in</strong>ormals. Provethat the curve is a geodesic on M.


§4. Covariant Differentiation, Parallel Translation, <strong>and</strong> Geodesics 7312. a. Prove that if a geodesic is planar <strong>and</strong> not a l<strong>in</strong>e, then it is a l<strong>in</strong>e of curvature. (H<strong>in</strong>t: UseExercise 11.)b. Prove that if every geodesic of a (connected) surface is planar, then the surface is conta<strong>in</strong>ed<strong>in</strong> a plane or a sphere.13. Prove or give a counterexample:a. A l<strong>in</strong>e ly<strong>in</strong>g <strong>in</strong> a surface is both an asymptotic curve <strong>and</strong> a geodesic.b. If a curve is both an asymptotic curve <strong>and</strong> a geodesic, then it must be a l<strong>in</strong>e.c. If a curve is both a geodesic <strong>and</strong> a l<strong>in</strong>e of curvature, then it must be planar.14. Show that the geodesic curvature at P of a curve C <strong>in</strong> M is equal (<strong>in</strong> absolute value) to thecurvature at P of the projection of C <strong>in</strong>to T P M.15. Check us<strong>in</strong>g Clairaut’s relation, Proposition 4.4, that great circles are geodesics on a sphere.(H<strong>in</strong>t: The result of Exercise A.1.3 may be useful.)16. Let M be asurface <strong>and</strong> P ∈ M. WesayU, V ∈ T P M are conjugate if II P (U, V) =0.a. Let C ⊂ M be a curve. Def<strong>in</strong>e the envelope M ∗ of the tangent planes to M along C tobe the ruled surface whose generator at P ∈ C is the limit<strong>in</strong>g position as Q → P of the<strong>in</strong>tersection l<strong>in</strong>e of the tangent planes to M at P <strong>and</strong> Q. Prove that the generator at P isconjugate to the tangent l<strong>in</strong>e to C at P .b. Prove that if C is nowhere tangent to an asymptotic direction, then M ∗ is smooth (at leastnear C). Prove, moreover, that M ∗ is tangent to M along C <strong>and</strong> is a developable (flatruled) surface.c. Apply part b to give a geometric way of comput<strong>in</strong>g parallel translation. In particular, dothis for a latitude circle on the sphere. (Cf. Exercise 7.)17. Suppose that on a surface M the parallel translation of a vector from one po<strong>in</strong>t to anotheris <strong>in</strong>dependent of the path chosen. Prove that M must be flat. (H<strong>in</strong>t: Fix an orthonormalbasis e o 1 , eo 2 for T P M <strong>and</strong> def<strong>in</strong>e vector fields e 1 , e 2 by parallel translat<strong>in</strong>g. Choose coord<strong>in</strong>atesso that the u-curves are always tangent to e 1 <strong>and</strong> the v-curves are always tangent to e 2 . SeeExercise 10.)18. Use the Clairaut relation, Proposition 4.4, to describe the geodesics on the torus as parametrized<strong>in</strong> Example 1(c) of Section 1. (Start with a geodesic start<strong>in</strong>g at <strong>and</strong> mak<strong>in</strong>g angle φ 0 withthe outer parallel. Your description should dist<strong>in</strong>guish between the cases cos φ 0 ≤ a−ba+b <strong>and</strong>cos φ 0 > a−ba+b. Which geodesics never cross the outer parallel at all? Also, remember thatthrough each po<strong>in</strong>t there is a unique geodesic <strong>in</strong> each direction.)19. Use the proof of the Clairaut relation, Proposition 4.4, to show that a geodesic on a surface ofrevolution is given <strong>in</strong> terms of the st<strong>and</strong>ard parametrization <strong>in</strong> Example 7 of Section 2 by∫v = cduf(u) √ f(u) 2 − c 2 + const.


74 Chapter 2. <strong>Surfaces</strong>: Local TheoryNow deduce that <strong>in</strong> the case of a non-arclength parametrization we obta<strong>in</strong>∫ √fv = c′ (u) 2 + g ′ (u) 2f(u) √ du + const.f(u) 2 − c2 20. Use Exercise 19 to show that any geodesic on the paraboloid z = x 2 + y 2 that is not a meridian<strong>in</strong>tersects every meridian. (H<strong>in</strong>t: Show that it cannot approach a meridian asymptotically.)21. Use the Clairaut relation, Proposition 4.4, <strong>and</strong> Exercise 19 to describe the geodesics on thepseudosphere (see Example 6 of Section 2). Show, <strong>in</strong> particular, that the only geodesics thatare unbounded are the meridians.22. Consider the surface z = f(u, v). A curve α whose tangent vector projects to a scalar multipleof ∇f(u, v) ateachpo<strong>in</strong>t P =(u, v, f(u, v)) is a curve of steepest ascent (why?). Suppose sucha curve α is also a geodesic.a. Prove that the projection of α <strong>in</strong>to the uv-plane is a geodesic. (H<strong>in</strong>t: The vector tangentto the surface <strong>and</strong> orthogonal to α ′ is horizontal. Why? Compare the Darboux framesalong α <strong>and</strong> its projection.)b. Deduce that α is also a l<strong>in</strong>e of curvature. (H<strong>in</strong>t: See Exercise 12.)c. Show that if all the curves of steepest ascent are geodesics, then f satisfies the partialdifferential equationf u f v (f vv − f uu )+f uv (f 2 u − f 2 v )=0.(H<strong>in</strong>t: When are the <strong>in</strong>tegral curves of ∇f l<strong>in</strong>es?)d. Show that the level curves of f are parallel (see Exercise 1.2.22). (H<strong>in</strong>t: Show that ‖∇f‖is constant along level curves.)e. Give a characterization of the surfaces with the property that all curves of steepest ascentare geodesics.


CHAPTER 3<strong>Surfaces</strong>: Further TopicsThe first section is required read<strong>in</strong>g, but the rema<strong>in</strong><strong>in</strong>g sections of this chapter are <strong>in</strong>dependentof one another.1. Holonomy <strong>and</strong> the Gauss-Bonnet TheoremLet’s now pursue the discussion of parallel translation that we began <strong>in</strong> Chapter 2. Let M be asurface <strong>and</strong> α a closed curve <strong>in</strong> M. Webeg<strong>in</strong> by fix<strong>in</strong>g a smoothly-vary<strong>in</strong>g orthonormal basis e 1 , e 2(a so-called fram<strong>in</strong>g) for the tangent planes of M <strong>in</strong> an open set of M conta<strong>in</strong><strong>in</strong>g α, asshown <strong>in</strong>Figure 1.2 below. Now we make the follow<strong>in</strong>ge 2e 2e 2e 1 e 2e 1ee 12e 1e 1αFigure 1.1Def<strong>in</strong>ition. Let α be a closed curve <strong>in</strong> a surface M. The angle through which a vector turnsrelative to the given fram<strong>in</strong>g as we parallel translate it once around the curve α is called theholonomy 1 around α.For example, if we take a fram<strong>in</strong>g around α by us<strong>in</strong>g the unit tangent vectors to α as our vectorse 1 , then, by the def<strong>in</strong>ition of a geodesic, there there will be zero holonomy around a closed geodesic(why?). For another example, if we use the fram<strong>in</strong>g on (most of) the sphere given by the tangentsto the l<strong>in</strong>es of longitude <strong>and</strong> l<strong>in</strong>es of latitude, the computation <strong>in</strong> Example 3 of Section 4 of Chapter2 shows that the holonomy around a latitude circle u = u 0 of the unit sphere is −2π cos u 0 .To make this more precise, for ease of underst<strong>and</strong><strong>in</strong>g, let’s work <strong>in</strong> an orthogonal parametrization2 <strong>and</strong> def<strong>in</strong>e a fram<strong>in</strong>g by sett<strong>in</strong>ge 1 = x u√E<strong>and</strong> e 2 = x v√G.1 from holo-+-nomy, the study of the whole2 This is no limitation away from umbilic po<strong>in</strong>ts, as we can apply Theorem 3.3 of the Appendix <strong>and</strong> use coord<strong>in</strong>atescom<strong>in</strong>g from the pr<strong>in</strong>cipal directions.75


76 Chapter 3. <strong>Surfaces</strong>: Further TopicsS<strong>in</strong>ce (much as <strong>in</strong> the case of curves) e 1 <strong>and</strong> e 2 give an orthonormal basis for the tangent space ofour surface at each po<strong>in</strong>t, all the <strong>in</strong>tr<strong>in</strong>sic curvature <strong>in</strong>formation (such as given by the Christoffelsymbols) is encapsulated <strong>in</strong> know<strong>in</strong>g how e 1 twists towards e 2 as we move around the surface. Inparticular, if α(t) =x(u(t),v(t)), a ≤ t ≤ b, isaparametrized curve, we can setφ 12 (t) = d dt(e1 (u(t),v(t)) ) · e 2 (u(t),v(t)),which we may write more casually as e ′ 1 (t) · e 2(t), with the underst<strong>and</strong><strong>in</strong>g that everyth<strong>in</strong>g must bedone <strong>in</strong> terms of the parametrization. We emphasize that φ 12 depends <strong>in</strong> an essential way on theparametrized curve α. Perhaps it’s better, then, to writeφ 12 = ∇ α ′e 1 · e 2 .Note, moreover, that the proof of Proposition 4.2 of Chapter 2 shows that ∇ α ′e 2 · e 1 = −φ 12 <strong>and</strong>∇ α ′e 1 · e 1 = ∇ α ′e 2 · e 2 =0. (Why?)Remark. Although the notation seems cumbersome, it rem<strong>in</strong>ds us that φ 12 is measur<strong>in</strong>g howe 1 twists towards e 2 as we move along the curve α. This notation will fit <strong>in</strong> a more general context<strong>in</strong> Section 3.Suppose now that α is a closed curve <strong>and</strong> we are <strong>in</strong>terested <strong>in</strong> the holonomy around α. Ife 1 happens to be parallel along α, then the holonomy will, of course, be 0. If not, let’s considerX(t) tobethe parallel translation of e 1 along α(t) <strong>and</strong> write X(t) =cos ψ(t)e 1 + s<strong>in</strong> ψ(t)e 2 , tak<strong>in</strong>gψ(0) = 0. Then X is parallel along α if <strong>and</strong> only if0 = ∇ α ′X = ∇ α ′(cos ψe 1 + s<strong>in</strong> ψe 2 )= cos ψ∇ α ′e 1 + s<strong>in</strong> ψ∇ α ′e 2 +(− s<strong>in</strong> ψe 1 + cos ψe 2 )ψ ′= cos ψφ 12 e 2 − s<strong>in</strong> ψφ 12 e 1 +(− s<strong>in</strong> ψe 1 + cos ψe 2 )ψ ′=(φ 12 + ψ ′ )(− s<strong>in</strong> ψe 1 + cos ψe 2 ).Thus, X is parallel along α if <strong>and</strong> only if ψ ′ (t) =−φ 12 (t). We therefore conclude:Proposition 1.1. The holonomy around the closed curve C equals ∆ψ = −∫ baφ 12 (t)dt.Example 1. Back to our example of the latitude circle u = u 0 on the unit sphere. Thene 1 = x u <strong>and</strong> e 2 =(1/ s<strong>in</strong> u)x v . Ifweparametrize the curve by tak<strong>in</strong>g t = v, 0≤ t ≤ 2π, then wehave (see Example 1 of Chapter 2, Section 3)∇ α ′e 1 = ∇ α ′x u = x uv = cot u 0 x v = cos u 0 e 2 ,<strong>and</strong> so φ 12 = cos u 0 . Therefore, the holonomy around the latitude circle (oriented counterclockwise)is ∆ψ = −∫ 2π0cos u 0 dt = −2π cos u 0 , confirm<strong>in</strong>g our previous results.Note that if we wish to parametrize the curve by arclength (as will be important shortly),we take s =(s<strong>in</strong> u 0 )v, 0≤ s ≤ 2π s<strong>in</strong> u 0 . Then, with respect to this parametrization, we haveφ 12 (s) =cot u 0 . (Why?) ▽


§1. Holonomy <strong>and</strong> the Gauss-Bonnet Theorem 77e 2α ′θe 1αFigure 1.2Suppose now that α is an arclength-parametrized curve <strong>and</strong> let’s write α(s) =x(u(s),v(s)) <strong>and</strong>T(s) =α ′ (s) =cos θ(s)e 1 + s<strong>in</strong> θ(s)e 2 , s ∈ [0,L], for a C 1 function θ(s) (cf. Lemma 3.6 of Chapter1), as <strong>in</strong>dicated <strong>in</strong> Figure 1.2. A formula fundamental for the rest of our work is the follow<strong>in</strong>g:Proposition 1.2. When α is an arclength-parametrized curve, the geodesic curvature of α isgiven byκ g (s) =φ 12 (s)+θ ′ 1(s) =2 √ (−Ev u ′ (s)+G u v ′ (s) ) + θ ′ (s).EGProof. Recall that κ g = κN · (n × T) =T ′ · (n × T). Now, s<strong>in</strong>ce T = cos θe 1 + s<strong>in</strong> θe 2 ,n × T = − s<strong>in</strong> θe 1 + cos θe 2 (why?), <strong>and</strong> soκ g = ∇ T T · (− s<strong>in</strong> θe 1 + cos θe 2 )= ∇ T (cos θe 1 + s<strong>in</strong> θe 2 ) · (− s<strong>in</strong> θe 1 + cos θe 2 )= ( cos θ∇ T e 1 + s<strong>in</strong> θ∇ T e 2)· (− s<strong>in</strong> θe1 + cos θe 2 )+ ( (− s<strong>in</strong> θ)θ ′ (− s<strong>in</strong> θ)+(cos θ)θ ′ (cos θ) )= (cos 2 θ + s<strong>in</strong> 2 θ)(φ 12 + θ ′ )=φ 12 + θ ′ ,as required.Next, we have (tak<strong>in</strong>g full advantage of the orthogonality of x u <strong>and</strong> x v )φ 12 = d ( )xu x√ · √ vds E Gby the formulas (‡) onp.55.= 1 √EG(xuu u ′ + x uv v ′) · x v= 1 √EG((Γuuu x u +Γ v uux v )u ′ +(Γ u uvx u +Γ v uvx v )v ′) · x v= G √EG(Γvuu u ′ +Γ v uvv ′) =□12 √ EG(−Ev u ′ + G u v ′) ,Remark. The first equality <strong>in</strong> Proposition 1.2 should not be surpris<strong>in</strong>g <strong>in</strong> the least. Curvatureof a plane curve measures the rate at which its unit tangent vector turns relative to a fixed referencedirection. Similarly, the geodesic curvature of a curve <strong>in</strong> a surface measures the rate at which itsunit tangent vector turns relative to a parallel vector field along the curve; θ ′ measures its turn<strong>in</strong>grelative to e 1 , which is itself turn<strong>in</strong>g at a rate given by φ 12 ,sothe geodesic curvature is the sum ofthose two rates.


78 Chapter 3. <strong>Surfaces</strong>: Further TopicsNow suppose that α is a closed curve bound<strong>in</strong>g a region R ⊂ M. Wedenote the boundary ofR by ∂R. Then by Green’s Theorem (see Theorem 2.6 of the Appendix), we have∫ L∫ L1φ 12 (s)ds =00 2 √ (−Ev u ′ (s)+G u v ′ (s) ) ∫1ds =EG∂R 2 √ (−Ev du + G u dv )EG∫∫ ( ( Ev) (=R 2 √ EG+ Gu) )v 2 √ dudvEG u(†)∫∫ (1 (=R 2 √ Ev) (√EG+ Gu) ) √EGdudv√EG v EG u } {{ }dA∫∫= − KdARby the formula (∗) for Gaussian curvature on p. 57. (Recall from the end of Section 1 of Chapter2 that the element of surface area on a parametrized surface is given by dA = ‖x u × x v ‖dudv =√EG − F 2 dudv.)We now see that Gaussian curvature <strong>and</strong> holonomy are <strong>in</strong>timately related:Corollary 1.3. When R is a region with smooth boundary ly<strong>in</strong>g <strong>in</strong> an orthogonal parametrization,the holonomy around ∂R is ∆ψ = ∫∫ R KdA.Proof. This follows immediately from Proposition 1.1 <strong>and</strong> the formula (†) above.□We conclude further from Proposition 1.2 that∫ ∫κ g ds = φ 12 ds + θ(L) − θ(0) ,∂R∂R } {{ }∆θso the total angle through which the tangent vector to ∂R turns is given by∫ ∫∫∆θ = κ g ds + KdA.∂RRNow, when R is simply connected (i.e., can be cont<strong>in</strong>uously deformed to a po<strong>in</strong>t), it is not toosurpris<strong>in</strong>g that ∆θ =2π. Intuitively, as we shr<strong>in</strong>k the curve to a po<strong>in</strong>t, e 1 becomes almost constantalong the curve, but the tangent vector must make one full rotation (as a consequence of theHopf Umlaufsatz, Theorem 3.5 of Chapter 1). S<strong>in</strong>ce ∆θ is an <strong>in</strong>tegral multiple of 2π that variescont<strong>in</strong>uously as we deform the curve, it must stay equal to 2π throughout.Corollary 1.4. If R is a simply connected region whose boundary curve is a geodesic, then∫∫KdA =∆θ =2π.RExample 2. We take R to be the upper hemisphere <strong>and</strong> use the usual spherical coord<strong>in</strong>atesparametrization. Then the unit tangent vector along ∂R is e 2 everywhere, so ∆θ =0,<strong>in</strong>contradictionwith Corollary 1.4. Alternatively, C = ∂R is a geodesic, so there should be zero holonomyaround C (computed with respect to this fram<strong>in</strong>g).How do we resolve this paradox? Well, although we’ve been sloppy about this po<strong>in</strong>t, thespherical coord<strong>in</strong>ates parametrization actually “fails” at the north pole (s<strong>in</strong>ce x v = 0). Indeed,there is no fram<strong>in</strong>g of the upper hemisphere with e 2 everywhere tangent to the equator. However,the reader can rest assured that there is some orthogonal parametrization of the upper hemisphere,


§1. Holonomy <strong>and</strong> the Gauss-Bonnet Theorem 79e.g., by stereographic projection from the south pole (cf. Example 1(e) <strong>in</strong> Section 1 of Chapter 2).▽Remark. In more advanced courses, the holonomy around the closed curve α is <strong>in</strong>terpretedas a rotation of the tangent plane of M at α(0). That is, what matters is ∆ψ (mod 2π), i.e.,the change <strong>in</strong> angle disregard<strong>in</strong>g multiples of 2π. This quantity does not depend on the choice offram<strong>in</strong>g e 1 , e 2 .We now set to work on one of the crown<strong>in</strong>g results of surface theory.Theorem 1.5 (Local Gauss-Bonnet). Suppose R is a simply connected region with piecewisesmooth boundary <strong>in</strong> a parametrized surface. If ∂R has exterior angles ɛ j , j =1,...,s, then∫ ∫∫s∑κ g ds + KdA+ ɛ j =2π.∂RRj=1ɛ2ɛ 1ɛ 3CFigure 1.3Note, as we <strong>in</strong>dicate <strong>in</strong> Figure 1.3, that we measure exterior angles so that |ɛ j |≤π for all j.Proof. If ∂R is smooth, then from our earlier discussion we <strong>in</strong>fer that∫ ∫∫κ g ds + KdA =∆θ =2π.∂RRBut the unit tangent vector turns less by the amount ∑ sj=1 ɛ j,sothe result follows. (We ask thereader to check the details <strong>in</strong> Exercise 11.) □Corollary 1.6. Forageodesic triangle (i.e., a region whose boundary consists of three geodesicsegments) R with <strong>in</strong>terior angles ι 1 , ι 2 , ι 3 ,wehave ∫∫ R KdA =(ι 1 + ι 2 + ι 3 ) − π, the angle excess.Proof. S<strong>in</strong>ce the boundary consists of geodesic segments, the geodesic curvature <strong>in</strong>tegral dropsout, <strong>and</strong> we are left with∫∫3∑3∑3∑KdA =2π − ɛ j =2π − (π − ι j )= ι j − π,Ras required.□j=1Remark. It is worthwhile to consider the three special cases K =0,K =1,K = −1, aspictured <strong>in</strong> Figure 1.4. When M is flat, the sum of the angles of a triangle is π, as<strong>in</strong>the Euclideancase. When M is positively curved, it takes more than π for the triangle to close up, <strong>and</strong> when Mj=1j=1


80 Chapter 3. <strong>Surfaces</strong>: Further TopicsK=0K=1 K=−1Figure 1.4is negatively curved, it takes less. Intuitively, this is because geodesics seem to “bow out” whenK>0 <strong>and</strong> “bow <strong>in</strong>” when K


§1. Holonomy <strong>and</strong> the Gauss-Bonnet Theorem 81nFigure 1.6(i) ∆ λ is the image of a triangle under an (orientation-preserv<strong>in</strong>g) orthogonal parametrization;(ii) ∆ λ ∩ ∆ µ is either empty, a s<strong>in</strong>gle vertex, or a s<strong>in</strong>gle edge;(iii) when ∆ λ ∩ ∆ µ consists of a s<strong>in</strong>gle edge, the orientations of the edge are opposite <strong>in</strong> ∆ λ<strong>and</strong> ∆ µ ; <strong>and</strong>(iv) at most one edge of ∆ λ is conta<strong>in</strong>ed <strong>in</strong> the boundary of M.We now make a st<strong>and</strong>ardDef<strong>in</strong>ition. Given a triangulation T of a surface M with V vertices, E edges, <strong>and</strong> F faces, wedef<strong>in</strong>e the Euler characteristic χ(M,T) =V − E + F .Example 4. We can triangulate a disk as shown <strong>in</strong> Figure 1.7, obta<strong>in</strong><strong>in</strong>g χ =1. Without∆ 2∆ 1∆ 3 ∆ 4V−E+F = 9−18+10 = 1V−E+F = 5−8+4 = 1Figure 1.7be<strong>in</strong>g so pedantic as to require that each ∆ λ be the image of a triangle under an orthogonalparametrization, we might just th<strong>in</strong>k of the disk as a s<strong>in</strong>gle triangle with its edges puffed out; thenwe would have χ = V − E + F =3− 3+1=1,aswell. We leave it to the reader to triangulate asphere <strong>and</strong> check that χ(Σ, T) =2. ▽The beautiful result to which we’ve been headed is now the follow<strong>in</strong>gTheorem 1.7 (Global Gauss-Bonnet). Let M be an oriented surface with piecewise-smoothboundary, equipped with a triangulation T as above. If ɛ k , k =1,...,l, are the exterior angles of∂M, then∫∂M∫∫κ g ds + KdA+Ml∑ɛ k =2πχ(M,T).Proof. As we illustrate <strong>in</strong> Figure 1.8, we will dist<strong>in</strong>guish vertices on the boundary <strong>and</strong> <strong>in</strong> the<strong>in</strong>terior, denot<strong>in</strong>g the respective total numbers by V b <strong>and</strong> V i . Similarly, we dist<strong>in</strong>guish among edgesk=1


82 Chapter 3. <strong>Surfaces</strong>: Further Topics<strong>in</strong>terior vertex<strong>in</strong>terior edge∆ µ∆ λFigure 1.8boundary edgesboundary vertex<strong>in</strong>terior/boundary edgeson the boundary, edges <strong>in</strong> the <strong>in</strong>terior, <strong>and</strong> edges that jo<strong>in</strong> a boundary vertex to an <strong>in</strong>terior vertex;we denote the respective numbers of these by E b , E i , <strong>and</strong> E ib .Now observe that∫∫m∑∫∫KdA = KdAR∆ λs<strong>in</strong>ce all the orientations are compatible, <strong>and</strong>∫κ g ds =∂Rλ=1m∑∫λ=1∂∆ λκ g dsbecause the l<strong>in</strong>e <strong>in</strong>tegrals over <strong>in</strong>terior <strong>and</strong> <strong>in</strong>terior/boundary edges cancel <strong>in</strong> pairs. Let ɛ λj , j =1, 2, 3, denote the exterior angles of the “triangle” ∆ λ . Then, apply<strong>in</strong>g Theorem 1.5 to ∆ λ ,wehave∫∫∫3∑κ g ds + KdA+ ɛ λj =2π,∂∆ λ ∆ λ<strong>and</strong> now, summ<strong>in</strong>g over the triangles, we obta<strong>in</strong>∫ ∫∫κ g ds + KdA+∂MMm∑λ=1 j=1j=13∑ɛ λj =2πm =2πF .Now we must do some careful account<strong>in</strong>g: Lett<strong>in</strong>g ι λj denote the respective <strong>in</strong>terior angles oftriangle ∆ λ ,wehave(∗)∑ɛ λj =∑(π − ι λj )=π(2E i + E ib ) − 2πV i<strong>in</strong>teriorvertices<strong>in</strong>teriorvertices<strong>in</strong>asmuch as each <strong>in</strong>terior edge contributes two <strong>in</strong>terior vertices, whereas each <strong>in</strong>terior/boundaryedge contributes just one, <strong>and</strong> the <strong>in</strong>terior angles at each <strong>in</strong>terior vertex sum to 2π. Next,(∗∗)∑boundaryverticesɛ λj = πE ib +l∑ɛ k .k=1


§1. Holonomy <strong>and</strong> the Gauss-Bonnet Theorem 83To see this, we reason as follows. Given a boundary vertex v, denote by a superscript (v) therelevant angle or number for which the vertex v is <strong>in</strong>volved. When v is a fixed smooth boundaryvertex, we haveπ = ∑ι (v)λ= ∑(π − ɛ (v)λ)=π(E(v) ib+1)− ∑ɛ (v)λ .v∈∆ λ v∈∆ λ v∈∆ λOn the other h<strong>and</strong>, when v is a fixed corner of ∂M with exterior angle ɛ k ,wehaveɛ k = π − ∑ι (v)λ= π − ∑(π − ɛ (v)λ)= ∑ɛ (v)λ− πE(v) ib .v∈∆ λ v∈∆ λ v∈∆ λThus,∑boundaryverticesɛ λj =∑v cornerɛ (v)λ + ∑v smoothɛ (v)λAdd<strong>in</strong>g equations (∗) <strong>and</strong> (∗∗) yields∑ɛ λj =λ,j∑<strong>in</strong>teriorverticesɛ λj +( l∑= ɛ k +k=1∑boundaryvertices∑v corner)πE (v)ib+ ∑πE (v)ib=v smoothɛ λj =2π(E i + E ib − V i )+l∑ɛ k .k=1l∑ɛ k + πE ib .At long last, therefore, our reckon<strong>in</strong>g concludes:∫ ∫∫l∑κ g ds + KdA+ ɛ k =2π ( )F − (E i + E ib )+V i =2π(V − E + F )=2πχ(M,T),∂MMk=1because we can deduce from the fact that the boundary curve ∂M is closed that V b = E b .We now derive some <strong>in</strong>terest<strong>in</strong>g conclusions:Corollary 1.8. The Euler characteristic χ(M,T) does not depend on the triangulation T ofM.Proof. The left-h<strong>and</strong>-side of the equality <strong>in</strong> Theorem 1.7 has noth<strong>in</strong>g whatsoever to do withthe triangulation. □It is therefore legitimate to denote the Euler characteristic by χ(M), with no reference to thetriangulation. It is proved <strong>in</strong> a course <strong>in</strong> algebraic topology that the Euler characteristic is a“topological <strong>in</strong>variant”; i.e., if we deform the surface M <strong>in</strong> a bijective, cont<strong>in</strong>uous manner (so as toobta<strong>in</strong> a homeomorphic surface), the Euler characteristic does not change. We therefore deduce:Corollary 1.9. The quantity∫∂M∫∫κ g ds + KdA+Ml∑ɛ kk=1is a topological <strong>in</strong>variant, i.e., does not change as we deform the surface M.In particular, <strong>in</strong> the event that ∂M = ∅, sothe surface M is closed, wehaveCorollary 1.10. When M is a closed, oriented surface without boundary, we have∫∫KdA =2πχ(M).Mk=1□


84 Chapter 3. <strong>Surfaces</strong>: Further TopicsIt is very <strong>in</strong>terest<strong>in</strong>g that the total curvature does not change as we deform the surface, for example,as shown <strong>in</strong> Figure 1.9. In a topology course, one proves that any closed, oriented surface without∫∫MKdA = 4πFigure 1.9boundary must have the topological type of a sphere or of a g-holed torus for some positive <strong>in</strong>tegerg. Thus (cf. Exercises 5 <strong>and</strong> 6), the possible Euler characteristics of such a surface are 2, 0, −2,−4, ...; moreover, the <strong>in</strong>tegral ∫∫ MKdA determ<strong>in</strong>es the topological type of the surface.We conclude this section with a few applications of the Gauss-Bonnet Theorem.Example 5. Suppose M is a surface of nonpositive Gaussian curvature. Then there cannot bea geodesic 2-gon R on M that bounds a simply connected region. For if there were, by Theorem1.5 we would have∫∫0 ≥ KdA =2π − (ɛ 1 + ɛ 2 ) > 0,Rwhich is a contradiction. (Note that the exterior angles must be strictly less than π because thereis a unique (smooth) geodesic with a given tangent direction.) ▽Example 6. Suppose M is topologically equivalent to a cyl<strong>in</strong>der <strong>and</strong> its Gaussian curvatureis negative. Then there is at most one simple closed geodesic <strong>in</strong> M. Note, first, as <strong>in</strong>dicated <strong>in</strong>Rα must be like one of theseFigure 1.10Figure 1.10, that if there is a simple closed geodesic α, itmust separate M (i.e., does not bounda region) or else it bounds a disk R, <strong>in</strong>which case we would have 0 > ∫∫ RKdA =2πχ(R) =2π,which is a contradiction. On the other h<strong>and</strong>, suppose there were two. If they don’t <strong>in</strong>tersect, thenthey bound a cyl<strong>in</strong>der R <strong>and</strong> we get 0 > ∫∫ RKdA =2πχ(R) =0,which is a contradiction. If theydo <strong>in</strong>tersect, then we we have a geodesic 2-gon bound<strong>in</strong>g a simply connected region, which cannothappen by Example 5. ▽


§1. Holonomy <strong>and</strong> the Gauss-Bonnet Theorem 85EXERCISES 3.11. Compute the holonomy around a parallel on*a. a torusb. the paraboloid x(u, v) =(u cos v, u s<strong>in</strong> v, u 2 )c. the catenoid x(u, v) =(cosh u cos v, cosh u s<strong>in</strong> v, u)2. Determ<strong>in</strong>e whether there can be a (smooth) closed geodesic on a surface whena. K>0b. K =0c. K0b. K =0c. K0 that is topologically a cyl<strong>in</strong>der. Prove that there cannot betwo disjo<strong>in</strong>t simple closed geodesics both go<strong>in</strong>g around the neck of the surface.9. Consider the paraboloid M parametrized by x(u, v) =(u cos v, u s<strong>in</strong> v, u 2 ), 0 ≤ u, 0≤ v ≤ 2π.Denote by M r that portion of the paraboloid def<strong>in</strong>ed by 0 ≤ u ≤ r. ∫a. Calculate the geodesic curvature of the boundary circle <strong>and</strong> compute κ g ds.∂M rb. Calculate χ(M r ).∫∫c. Use the Gauss-Bonnet Theorem to compute KdA. F<strong>in</strong>d the limit as r →∞. (ThisM ris the total curvature of the paraboloid.)∫∫d. Calculate K directly (however you wish) <strong>and</strong> compute KdA explicitly.e. Expla<strong>in</strong> the relation between the total curvature <strong>and</strong> the image of the Gauss map of M.10. Consider the pseudosphere (with boundary) M parametrized as <strong>in</strong> Example 6 of Chapter 2,Section 2, but here we take u ≥ 0. Denote by M r that portion def<strong>in</strong>ed by 0 ≤ u ≤ r. (Notethat we are <strong>in</strong>clud<strong>in</strong>g the boundary circle u = 0.)M


86 Chapter 3. <strong>Surfaces</strong>: Further Topicsa. Calculate the geodesic curvature of the circle u = u 0 <strong>and</strong> compute∫κ g ds.∂M rb. Calculate χ(M r ).∫∫c. Use the Gauss-Bonnet Theorem to compute KdA. F<strong>in</strong>d the limit as r →∞. (ThisM ris the total curvature of the pseudosphere.)∫∫d. Calculate the area of M r directly, <strong>and</strong> use this to deduce the value of KdA.e. Expla<strong>in</strong> the relation between the total curvature <strong>and</strong> the image of the Gauss map of M.11. Check the details of the proof of Theorem 1.5 <strong>in</strong> the presence of corners. (See Exercise 1.3.6.)12. a. Use Corollary 1.3 to prove that M is flat if <strong>and</strong> only if the holonomy around all (“small”)closed curves that bound a region <strong>in</strong> M is zero.b. Show that even on a flat surface, holonomy can be nontrivial around certa<strong>in</strong> curves.13. Reprove the result of Exercise 2.3.13 by consider<strong>in</strong>g the holonomy around a (sufficiently small)quadrilateral formed by four of the l<strong>in</strong>es. Does the result hold if there are two families ofgeodesics <strong>in</strong> M always <strong>in</strong>tersect<strong>in</strong>g at right angles?14. Suppose α is a closed space curve with κ ≠0. Assume that the normal <strong>in</strong>dicatrix (i.e., thecurve traced out on the unit sphere by the pr<strong>in</strong>cipal normal) is a simple closed curve <strong>in</strong> theunit sphere. Prove then that it divides the unit sphere <strong>in</strong>to two regions of equal area. (H<strong>in</strong>t:Apply the Gauss-Bonnet Theorem to one of those regions.)15. Suppose M ⊂ R 3 is a closed, oriented surface with no boundary with K>0. It follows thatM is topologically a sphere (why?). Prove that M is convex; i.e., for each P ∈ M, M lies ononly one side of the tangent plane T P M. (H<strong>in</strong>t: Use the Gauss-Bonnet Theorem <strong>and</strong> Gauss’sorig<strong>in</strong>al <strong>in</strong>terpretation of curvature <strong>in</strong>dicated <strong>in</strong> the remark on p. 48 to show the Gauss mapmust be one-to-one. Then look at the end of the proof of Theorem 3.4 of Chapter 1.)M2. An Introduction to Hyperbolic GeometryHilbert proved <strong>in</strong> 1901 that there is no surface (without boundary) <strong>in</strong> R 3 with constant negativecurvature with the property that it is a closed subset of R 3 (i.e., every Cauchy sequence of po<strong>in</strong>ts<strong>in</strong> the surface converges to a po<strong>in</strong>t of the surface). The pseudosphere fails the latter condition.Nevertheless, it is possible to give a def<strong>in</strong>ition of an “abstract surface” (not sitt<strong>in</strong>g <strong>in</strong>side R 3 )together with a first fundamental form. As we know, this will be all we need to calculate Christoffelsymbols, curvature (Theorem 3.1 of Chapter 2), geodesics, <strong>and</strong> so on.Def<strong>in</strong>ition. The hyperbolic plane H is def<strong>in</strong>ed to be the half-plane {(u, v) ∈ R 2 : v > 0},equipped with the first fundamental form I given by E = G =1/v 2 , F =0.Now, us<strong>in</strong>g the formulas (‡) onp.55, we f<strong>in</strong>d thatΓ u uu = E u2E =0Γv uu = − E v2G = 1 v


§2. An Introduction to Hyperbolic Geometry 87Γ u uv = E v2E = −1 vΓvv u = − G u2E =0Us<strong>in</strong>g the formula (∗) for Gaussian curvature on p. 57, we f<strong>in</strong>dK = − 1 ( (2 √ Ev) (√EG+ Gu) )√ = − v2EG v EG u 2Γ v uv = G u2G =0Γv vv = G v2G = −1 v .(− 2 v 3 · v2) v = −v2 2 ·2v 2 = −1.Thus, the hyperbolic plane has constant curvature −1. Note that it is a consequence of Corollary1.6 that the area of a geodesic triangle <strong>in</strong> H is equal to π − (ι 1 + ι 2 + ι 3 ).What are the geodesics <strong>in</strong> this surface? Us<strong>in</strong>g the equations (♣♣) onp.68, we obta<strong>in</strong> theequationsu ′′ − 2 v u′ v ′ = v ′′ + 1 v (u′2 − v ′2 )=0.Obviously, the vertical rays u = const give us solutions (with v(t) =c 1 e c2t ). Next we seek geodesicswith u ′ ≠0,sowelook for v as a function of u <strong>and</strong> obta<strong>in</strong> (with very careful use of the cha<strong>in</strong> rule)d 2 vdu 2 = d ( ) v′du u ′ = u′ v ′′ − u ′′ v ′u ′2 · 1u ′= 1 ) (v ′2 − u ′2) ( )) 2− v ′ v u′ v ′( 1(u ′u ′3 v(= − 1 ( ) ) (v′ 21+v u ′ = − 1 1+vThis means we are left with the differential equation( )v d2 v dv 2du 2 + = d (v dv )= −1,du du du<strong>and</strong> <strong>in</strong>tegrat<strong>in</strong>g this twice gives us the solutionsu 2 + v 2 = au + b.( ) )dv2.duThat is, the geodesics <strong>in</strong> H are the vertical rays <strong>and</strong> the semicircles centered on the u-axis, aspictured <strong>in</strong> Figure 2.1. Note that any semicircle centered on the u-axis <strong>in</strong>tersects each vertical l<strong>in</strong>evuFigure 2.1at most one time. It now follows that any two po<strong>in</strong>ts P, Q ∈ H are jo<strong>in</strong>ed by a unique geodesic. If P


88 Chapter 3. <strong>Surfaces</strong>: Further Topics<strong>and</strong> Q lie on a vertical l<strong>in</strong>e, then the vertical ray through them is the unique geodesic jo<strong>in</strong><strong>in</strong>g them.If P <strong>and</strong> Q do not lie on a vertical l<strong>in</strong>e, let C be the <strong>in</strong>tersection of the perpendicular bisector ofPQ <strong>and</strong> the u-axis; then the semicircle centered at C is the unique geodesic jo<strong>in</strong><strong>in</strong>g P <strong>and</strong> Q.Example 1. Given P, Q ∈ H, wewould like to f<strong>in</strong>d a formula for the (geodesic) distanced(P, Q) between them. Let’s start with P =(u 0 ,a) <strong>and</strong> Q =(u 0 ,b), with 0


§2. An Introduction to Hyperbolic Geometry 89[ ] a b(i) Composition of functions corresponds to multiplication of the 2 × 2 matriceswith determ<strong>in</strong>ant 1, so we obta<strong>in</strong> a group.(ii) T maps H bijectively to H.(iii) T is an isometry of H.We leave it to the reader to check the first two <strong>in</strong> Exercise 6, <strong>and</strong> we check the third here. Giventhe po<strong>in</strong>t z = u + iv, wewant to compute the lengths of the vectors T u <strong>and</strong> T v at the image po<strong>in</strong>tT (z) =x + iy <strong>and</strong> see that the two vectors are orthogonal. Note thatso y =so we have(a(u + iv)+b)(c(u − iv)+d)az + b (az + b)(cz + d)=cz + d |cz + d| 2 =|cz + d| 2(ac(u 2 + v 2 )+(ad + bc)u ) + i ( (ad − bc)v )=v|cz + d| 2 .Nowwehave3<strong>and</strong>, similarly, ˜G =x 2 v + y 2 vy 2x u + iy u = −ix v + y v = T ′ (z) =|cz + d| 2 ,(cz + d)a − (az + b)c(cz + d) 2 =Ẽ = x2 u + yu2y 2 = 1 y 2 |T ′ (z)| 2 = 1 y 2 · 1|cz + d| 4 = 1 v 2 = E,= G. Onthe other h<strong>and</strong>,˜F = x uy u + x v y vy 2 = x u(−x v )+x v (x u )y 2 =0=F,1(cz + d) 2 ,as desired.Now, as we verify <strong>in</strong> Exercise 10 or <strong>in</strong> Exercise 12, l<strong>in</strong>ear fractional transformations carryl<strong>in</strong>es <strong>and</strong> circles <strong>in</strong> C to either l<strong>in</strong>es or circles. S<strong>in</strong>ce our particular l<strong>in</strong>ear fractional transformationspreserve the real axis (∪{∞}) <strong>and</strong> preserve angles as well, it follows that vertical l<strong>in</strong>es <strong>and</strong> semicirclescentered on the real axis map to one another. Thus, our isometries do <strong>in</strong> fact map geodesics togeodesics (how comfort<strong>in</strong>g!).If we th<strong>in</strong>k of H as model<strong>in</strong>g non-Euclidean geometry, with l<strong>in</strong>es <strong>in</strong> our geometry be<strong>in</strong>g thegeodesics, note that given any l<strong>in</strong>e l <strong>and</strong> po<strong>in</strong>t P /∈ l, there are <strong>in</strong>f<strong>in</strong>itely many l<strong>in</strong>es pass<strong>in</strong>gthrough P “parallel” to (i.e., not <strong>in</strong>tersect<strong>in</strong>g) l. Aswesee <strong>in</strong> Figure 2.3, there are also two specialcdlPFigure 2.3l<strong>in</strong>es through P that “meet l at <strong>in</strong>f<strong>in</strong>ity”; these are called ultraparallels.3 These are the Cauchy-Riemann equations from basic complex analysis.


90 Chapter 3. <strong>Surfaces</strong>: Further TopicsWe conclude with an <strong>in</strong>terest<strong>in</strong>g application. As we saw <strong>in</strong> the previous section, the Gauss-Bonnet Theorem gives a deep relation between the total curvature of a surface <strong>and</strong> its topologicalstructure (Euler characteristic). We know that if a compact surface M is topologically equivalent toa sphere, then its total curvature must be that of a round sphere, namely 4π. IfM is topologicallyequivalent to a torus, then (as the reader checked <strong>in</strong> Exercise 3.1.4) its total curvature must be0. We know that there is no way of mak<strong>in</strong>g the torus <strong>in</strong> R 3 <strong>in</strong> such a way that it has constantFigure 2.4Gaussian curvature K =0(why?), but we can construct a flat torus <strong>in</strong> R 4 by tak<strong>in</strong>gx(u, v) =(cos u, s<strong>in</strong> u, cos v, s<strong>in</strong> v), 0 ≤ u, v ≤ 2π.(We take a piece of paper <strong>and</strong> identify opposite edges, as <strong>in</strong>dicated <strong>in</strong> Figure 2.4; this can be rolled<strong>in</strong>to a cyl<strong>in</strong>der <strong>in</strong> R 3 but <strong>in</strong>to a torus only <strong>in</strong> R 4 .) So what happens with a 2-holed torus? Inthat case, χ(M) =−2, so the total curvature should be −4π, <strong>and</strong> we can reasonably ask if there’sa 2-holed torus with constant negative curvature. Note that we can obta<strong>in</strong> a 2-holed torus byidentify<strong>in</strong>g pairs of edges on an octagon, as shown <strong>in</strong> Figure 2.5.dabcadcbFigure 2.5This leads us to wonder whether we might have regular n-gons R <strong>in</strong> H. Bythe Gauss-Bonnetformula, we would have area(R) =(n−2)π − ∑ ι j ,soit’s obviously necessary that ∑ ι j < (n−2)π.This shouldn’t be difficult so long as n ≥ 3. <strong>First</strong>, let’s conv<strong>in</strong>ce ourselves that, given any po<strong>in</strong>tP ∈ H, 0


§2. An Introduction to Hyperbolic Geometry 91PαrβQrRFigure 2.6(divid<strong>in</strong>g the angle at P <strong>in</strong>to n angles of 2π/n each), we obta<strong>in</strong> a regular n-gon with the propertythat ∑ ι j =2π, asshown (approximately?) <strong>in</strong> Figure 2.7 for the case n =8. The po<strong>in</strong>t is thatFigure 2.7because the <strong>in</strong>terior angles add up to 2π, when we identify edges as <strong>in</strong> Figure 2.5, we will obta<strong>in</strong> asmooth 2-holed torus with constant curvature K = −1. The analogous construction works for theg-holed torus, construct<strong>in</strong>g a regular 4g-gon whose <strong>in</strong>terior angles sum to 2π.EXERCISES 3.21. F<strong>in</strong>d the geodesic jo<strong>in</strong><strong>in</strong>g P <strong>and</strong> Q <strong>in</strong> H <strong>and</strong> calculate d(P, Q).*a. P =(0, 1), Q =(1, 2)b. P =(−1, 8), Q =(2, 7)c. P =(0, 9), Q =(4, 7)2. Suppose there is a geodesic perpendicular to two geodesics <strong>in</strong> H. What can you prove aboutthe latter two?3. Prove the angle-angle-angle congruence theorem for hyperbolic triangles: If ∠A ∼ = ∠A ′ , ∠B ∼ =∠B ′ , <strong>and</strong> ∠C ∼ = ∠C ′ , then △ABC ∼ = △A ′ B ′ C ′ . (H<strong>in</strong>t: Use an isometry to move A ′ to A, B ′along the geodesic from A to B, <strong>and</strong> C ′ along the geodesic from A to C.)


92 Chapter 3. <strong>Surfaces</strong>: Further Topics*4. Us<strong>in</strong>g Proposition 1.2, prove that horizontal l<strong>in</strong>es have constant geodesic curvature κ g =1<strong>in</strong>H; prove that this is also true for circles tangent to the u-axis. (These curves are classicallycalled horocycles.)5. Prove that every curve <strong>in</strong> H of constant geodesic curvature κ g =1is of the form given <strong>in</strong>Exercise 4. (H<strong>in</strong>ts: Assume we start with an arclength parametrization (u(s),v(s)), <strong>and</strong> useProposition 1.2 to show that we have 1 = u′v + θ′ <strong>and</strong> u ′2 + v ′2 = v 2 . Obta<strong>in</strong> the differentialequation(v d2 v( dv) ) 2 3/2 ( dv) 2du 2 = 1+− − 1,dudu<strong>and</strong> solve this by substitut<strong>in</strong>g z = dv/du <strong>and</strong> gett<strong>in</strong>g a separable differential equation fordz/dv.)6. Let T a,b,c,d (z) = az + b , a, b, c, d ∈ R, with ad − bc =1.cz + da. Suppose a ′ ,b ′ ,c ′ ,d ′ ∈ R <strong>and</strong> a ′ d ′ − b ′ c ′ =1. Check thatT a ′ ,b ′ ,c ′ ,d ′◦ T a,b,c,d = T a ′ a+b ′ c,a ′ b+b ′ d,c ′ a+d ′ c,c ′ b+d ′ d(a ′ a + b ′ c)(c ′ b + d ′ d) − (a ′ b + b ′ d)(c ′ a + d ′ c)=1.Show, moreover, that T d,−b,−c,a = T −1a,b,c,d . (Note that T a,b,c,d = T −a,−b,−c,−d . The readerwho’s taken group theory will recognize that we’re def<strong>in</strong><strong>in</strong>g an isomorphism between thegroup of l<strong>in</strong>ear fractional transformations <strong>and</strong> the group SL(2, R)/{±I} of 2 × 2 matriceswith determ<strong>in</strong>ant 1, identify<strong>in</strong>g a matrix <strong>and</strong> its additive <strong>in</strong>verse.)b. Let T = T a,b,c,d . Prove that if z = u + iv <strong>and</strong> v>0, then T (z) =x + iy with y>0.Deduce that T maps H to itself bijectively.7. Show that reflection across the geodesic u =0isgiven by r(z) =−z. Use this to determ<strong>in</strong>ethe form of the reflection across a general geodesic.8. The geodesic circle of radius R centered at P is the set of po<strong>in</strong>ts Q so that d(P, Q) =R. Provethat geodesic circles <strong>in</strong> H are Euclidean circles. One way to proceed is as follows: The geodesiccircle centered at P =(0, 1) with radius R =lna must pass through (0,a) <strong>and</strong> (0, 1/a), <strong>and</strong>hence ought to be a Euclidean circle centered at (0, 1 2(a +1/a)). Check that all the po<strong>in</strong>ts onthis circle are <strong>in</strong> fact a hyperbolic distance R away from P . (H<strong>in</strong>t: It is probably easiest towork with the cartesian equation of the circle.)Remark. This result can be established rather pa<strong>in</strong>lessly by us<strong>in</strong>g the Po<strong>in</strong>caré disk modelgiven <strong>in</strong> Exercise 15.<strong>and</strong>*9. What is the geodesic curvature of a geodesic circle of radius R <strong>in</strong> H? (See Exercises 4 <strong>and</strong> 8.)10. Recall (see, for example, p. 298 <strong>and</strong> pp. 350–1 of Shifr<strong>in</strong>’s Abstract Algebra: A GeometricApproach) that the cross ratio of four numbers A, B, P, Q ∈ C ∪ {∞} is def<strong>in</strong>ed to be[A : B : P : Q] = Q − A / Q − BP − A P − B .


§2. An Introduction to Hyperbolic Geometry 93a. Show that A, B, P , <strong>and</strong> Q lie on a l<strong>in</strong>e or circle if <strong>and</strong> only if their cross ratio is a realnumber.b. Prove that if S is a l<strong>in</strong>ear fractional transformation with S(A) = 0, S(B) = ∞, <strong>and</strong>S(P )=1, then S(Q) =[A : B : P : Q]. Use this to deduce that for any l<strong>in</strong>ear fractionaltransformation T ,wehave[T (A) :T (B) :T (P ):T (Q)] = [A : B : P : Q].c. Prove that l<strong>in</strong>ear fractional transformations map l<strong>in</strong>es <strong>and</strong> circles to either l<strong>in</strong>es or circles.(For which such transformations do l<strong>in</strong>es necessarily map to l<strong>in</strong>es?)d. Show that if A, B, P , <strong>and</strong> Q lie on a l<strong>in</strong>e or circle, then[A : B : P : Q] = AQ / BQAP BP .Conclude that d(P, Q) =| ln[A : B : P : Q]|, where A, B, P , <strong>and</strong> Q are as illustrated <strong>in</strong>Figure 2.2.e. Check that if T is a l<strong>in</strong>ear fractional transformation carry<strong>in</strong>g A to 0, B to ∞, P to P ′ ,<strong>and</strong> Q to Q ′ , then d(P, Q) =d(P ′ ,Q ′ ).11. a. Let O be any po<strong>in</strong>t not ly<strong>in</strong>g on a circle C <strong>and</strong> let P <strong>and</strong> Q be po<strong>in</strong>ts on the circle C sothat O, P , <strong>and</strong> Q are coll<strong>in</strong>ear. Let T be the po<strong>in</strong>t on C so that OT is tangent to C. Provethat (OP)(OQ) =(OT) 2 .b. Def<strong>in</strong>e <strong>in</strong>version <strong>in</strong> the circle of radius R centered at O by send<strong>in</strong>g a po<strong>in</strong>t P to the po<strong>in</strong>tP ′ on the ray OP with (OP)(OP ′ )=R 2 . Show that an <strong>in</strong>version <strong>in</strong> a circle centered atthe orig<strong>in</strong> maps a circle C centered on the u-axis <strong>and</strong> not pass<strong>in</strong>g through O to anothercircle C ′ centered on the u-axis. (H<strong>in</strong>t: For any P ∈ C, let Q be the other po<strong>in</strong>t on Ccoll<strong>in</strong>ear with O <strong>and</strong> P , <strong>and</strong> let Q ′ be the image of Q under <strong>in</strong>version. Use the result ofpart a to show that OP/OQ ′ is constant. If C is the center of C, let C ′ be the po<strong>in</strong>t onthe u-axis so that C ′ Q ′ ‖CP. Show that Q ′ traces out a circle C ′ centered at C ′ .)c. Show that <strong>in</strong>version <strong>in</strong> the circle of radius R centered at O maps vertical l<strong>in</strong>es to circlescentered on the u-axis <strong>and</strong> pass<strong>in</strong>g through O <strong>and</strong> vice-versa.12. a. Prove that every (orientation-preserv<strong>in</strong>g) isometry of H can be written as the compositionof l<strong>in</strong>ear fractional transformations of the formT 1 (z) =z + b for some b ∈ R <strong>and</strong> T 2 (z) =− 1 z .(H<strong>in</strong>t:[It’s]probably[ ]easiest to work[with]matrices. Show that you have matrices of the1 b 0 −11 0form , , <strong>and</strong> therefore , <strong>and</strong> that any matrix of determ<strong>in</strong>ant 1 can be0 1 1 0b 1obta<strong>in</strong>ed by row <strong>and</strong> column operations from these.)b. Prove that T 2 maps circles centered on the u-axis <strong>and</strong> vertical l<strong>in</strong>es to circles centered onthe u-axis <strong>and</strong> vertical l<strong>in</strong>es (not necessarily respectively). Either do this algebraically oruse Exercise 11.c. Use the results of parts a <strong>and</strong> b to prove that isometries of H map geodesics to geodesics.13. We say a l<strong>in</strong>ear fractional transformation T = T a,b,c,d is elliptic if it has one fixed po<strong>in</strong>t, parabolicif it has one fixed po<strong>in</strong>t at <strong>in</strong>f<strong>in</strong>ity, <strong>and</strong> hyperbolic if it has two fixed po<strong>in</strong>ts at <strong>in</strong>f<strong>in</strong>ity.a. Show that T is elliptic if |a + d| < 2, parabolic if |a + d| =2,<strong>and</strong> hyperbolic if |a + d| > 2.


94 Chapter 3. <strong>Surfaces</strong>: Further Topicsb. Describe the three types of isometries geometrically. (H<strong>in</strong>t: In particular, what is therelation between horocycles <strong>and</strong> parabolic l<strong>in</strong>ear fractional transformations?)14. Suppose △ABC is a hyperbolic right triangle with “hypotenuse” c. Use Figure 2.8 to provethe follow<strong>in</strong>g:s<strong>in</strong> ∠A = s<strong>in</strong>h as<strong>in</strong>h c ,tanh bcos ∠A = , cosh c = cosh a cosh b.tanh b(The last is the hyperbolic Pythagorean Theorem.) (H<strong>in</strong>t: Start by show<strong>in</strong>g, for example, thatRψτBa cC brθAFigure 2.8cosh b = csc θ, cosh c =(1− cos ψ cos τ)/(s<strong>in</strong> ψ s<strong>in</strong> τ), <strong>and</strong> cos τ − cos ψ = s<strong>in</strong> τ cot θ. You willneed two equations trigonometrically relat<strong>in</strong>g R <strong>and</strong> r.)15. Given a po<strong>in</strong>t P on a surface M, wedef<strong>in</strong>e the geodesic circle of radius R centered at P to bethe locus of po<strong>in</strong>ts whose (geodesic) distance from P is R. Let C(R) denote its circumference.a. Show that on the unit sphere2πR − C(R)limR→0 + πR 3 = 1 3 .b. Show that the geodesic curvature κ g of a spherical geodesic circle of radius R iscot R ≈ 1 R2R(1 −3 + ...).The Po<strong>in</strong>caré disk is def<strong>in</strong>ed to be the “abstract surface” D = {(u, v) :u 2 + v 2 < 1} with the4first fundamental form given by E = G =(1 − u 2 − v 2 , F =0. This is called the hyperbolic) 2metric on D.c. Check that <strong>in</strong> D the geodesics through the orig<strong>in</strong> are Euclidean l<strong>in</strong>e segments; concludethat the ( Euclidean ) circle of radius r centered at the orig<strong>in</strong> is a hyperbolic circle of radius1+rR =ln , <strong>and</strong> so r = tanh R 1 − r2 .d. Check that the circumference of the hyperbolic circle is 2π s<strong>in</strong>h R ≈2π(R + R36 + ...), <strong>and</strong> so 2πR − C(R)limR→0 + πR 3 = − 1 3 .e. Compute (us<strong>in</strong>g a double <strong>in</strong>tegral) that the area of a disk of hyperbolic radius R is4π s<strong>in</strong>h 2 R 2≈ πR2 (1 + R212+ ...). Use the Gauss-Bonnet theorem to deduce that thegeodesic curvature κ g of the hyperbolic circle of radius R is coth R ≈ 1 R2R(1 +3 + ...).16. Here we give another model for hyperbolic geometry, called the Kle<strong>in</strong>-Beltrami model. Considerthe follow<strong>in</strong>g parametrization of the hyperbolic disk: Start with the open unit disk,


§3. Surface Theory with Differential Forms 95{x21 + x 2 2 < 1, x 3 =0 } ,vertically project to the southern hemisphere of the unit sphere, <strong>and</strong>then stereographically project (from the north pole) back to the unit disk.a. Show that this mapp<strong>in</strong>g is given (<strong>in</strong> terms of polar coord<strong>in</strong>ates) by()Rx(R, θ) =(r, θ) =1+ √ 1 − R ,θ .2Compute that the first fundamental form of the Po<strong>in</strong>caré metric on D (see Exercise 15),4E =(1 − r 2 ) 2 , F =0,G = 4r 2(1 − r 2 ) 2 ,isgiven <strong>in</strong> (R, θ) coord<strong>in</strong>ates by Ẽ = 1(1 − R 2 ) 2 ,R˜F =0, ˜G 2= . (H<strong>in</strong>t: Compute carefully <strong>and</strong> economically!)1 − R2 b. Chang<strong>in</strong>g now to Euclidean coord<strong>in</strong>ates (u, v), show thatÊ =whence you derive1 − v 2(1 − u 2 − v 2 ) 2 , ˆF =uv(1 − u 2 − v 2 ) 2 , Ĝ =Γuu u 2u=1 − u 2 − v 2 , Γ uu v =0,Γuv u v=1 − u 2 − v 2 , Γ uv v u=1 − u 2 − v 2 ,Γ uvv =0,Γ vvv =1 − u 2(1 − u 2 − v 2 ) 2 ,2v1 − u 2 − v 2 .c. Use part b to show that the geodesics of the disk us<strong>in</strong>g the first fundamental form Î arechords of the circle u 2 + v 2 =1. (H<strong>in</strong>t: Show (by us<strong>in</strong>g the cha<strong>in</strong> rule) that the equationsfor a geodesic give d 2 v/du 2 = 0.) Discuss the advantages <strong>and</strong> disadvantages of this model(compared to Po<strong>in</strong>caré’s).d. Compute the distance from (0, 0) to (a, 0); compare with the formula for distance <strong>in</strong> thePo<strong>in</strong>caré model.e. Check your answer <strong>in</strong> part c by prov<strong>in</strong>g (geometrically?) that chords of the circle map byx to geodesics <strong>in</strong> the hyperbolic disk. (See Exercise 2.1.6.)3. Surface Theory with Differential FormsWe’ve seen that it can be quite awkward to work with coord<strong>in</strong>ates to study surfaces. (Forexample, the Codazzi <strong>and</strong> Gauss Equations <strong>in</strong> Section 3 of Chapter 2 are far from beautiful.) Forthose who’ve learned about differential forms, we can given a quick <strong>and</strong> elegant treatment that isconceptually quite clean.We start (much like the situation with curves) with a mov<strong>in</strong>g frame e 1 , e 2 , e 3 on (an open subsetof) our (oriented) surface M. Here e i are vector fields def<strong>in</strong>ed on M with the properties that(i) {e 1 , e 2 , e 3 } gives an orthonormal basis for R 3 at each po<strong>in</strong>t (so the matrix with thoserespective column vectors is an orthogonal matrix);(ii) {e 1 , e 2 } is a basis for the tangent space of M <strong>and</strong> e 3 = n.


96 Chapter 3. <strong>Surfaces</strong>: Further TopicsHow do we know such a mov<strong>in</strong>g frame exists? If x: U → M is a parametrized surface, we can startwith our usual vectors x u , x v <strong>and</strong> apply the Gram-Schmidt process to obta<strong>in</strong> an orthonormal basis.Or, if U is a region conta<strong>in</strong><strong>in</strong>g no umbilic po<strong>in</strong>ts, then we can choose e 1 <strong>and</strong> e 2 to be unit vectorspo<strong>in</strong>t<strong>in</strong>g <strong>in</strong> the pr<strong>in</strong>cipal directions (this approach was tacit <strong>in</strong> many of our proofs earlier).If x: M → R 3 is the <strong>in</strong>clusion map (which we may choose, <strong>in</strong> a computational sett<strong>in</strong>g, toconsider as the parametrization mapp<strong>in</strong>g U → R 3 ), then we def<strong>in</strong>e 1-forms ω 1 ,ω 2 on M bydx = ω 1 e 1 + ω 2 e 2 ;i.e., for any V ∈ T P M,wehave V = ω 1 (V)e 1 + ω 2 (V)e 2 ,soω α (V) =I(V, e α ) for α =1, 2. So far,ω 1 <strong>and</strong> ω 2 keep track of how our po<strong>in</strong>t moves around on M. Next we want to see how the frameitself twists, so we def<strong>in</strong>e 1-forms ω ij , i, j =1, 2, 3, byde i =3∑ω ij e j .j=1Note that s<strong>in</strong>ce e i · e j = const for any i, j =1, 2, 3, we have( 3∑ )0=d(e i · e j )=de i · e j + e i · de j = ω ik e k · e j += ω ij + ω ji ,k=1( 3∑k=1ω jk e k)· e iso ω ji = −ω ij for all i, j =1, 2, 3. (In particular, s<strong>in</strong>ce e i is always a unit vector, ω ii =0for all i.)If V ∈ T P M, ω ij (V) tells us how fast e i is twist<strong>in</strong>g towards e j at P as we move with velocity V.Note, <strong>in</strong> particular, that the shape operator is embodied <strong>in</strong> the equationde 3 = ω 31 e 1 + ω 32 e 2 = − ( ω 13 e 1 + ω 23 e 2).Then for any V ∈ T P M we have ω 13 (V) =II(V, e 1 ) <strong>and</strong> ω 23 (V) =II(V, e 2 ). Indeed, when wewriteω 13 = h 11 ω 1 + h 12 ω 2ω 23 = h 21 ω 1 + h 22 ω 2for appropriate coefficient functions h αβ ,wesee that the matrix of the shape operator S Prespect to the basis {e 1 , e 2 } for T P M is noth<strong>in</strong>g but [ ]h αβ .Most of our results will come from the follow<strong>in</strong>gwithTheorem 3.1 (Structure Equations).dω 1 = ω 12 ∧ ω 2 <strong>and</strong> dω 2 = ω 1 ∧ ω 12 , <strong>and</strong>3∑dω ij = ω ik ∧ ω kj for all i, j =1, 2, 3.k=1Proof. From the properties of the exterior derivative, we have( 3∑ ) ( 3∑ )0 = d(dx) =dω 1 e 1 + dω 2 e 2 − ω 1 ∧ ω 1j e j − ω 2 ∧ ω 2j e jj=1j=1= ( dω 1 − ω 2 ∧ ω 21)e1 + ( dω 2 − ω 1 ∧ ω 12)e2 − ( ω 1 ∧ ω 13 + ω 2 ∧ ω 23)e3 ,


§3. Surface Theory with Differential Forms 97so from the fact that {e 1 , e 2 , e 3 } is a basis for R 3 we <strong>in</strong>fer thatSimilarly, we obta<strong>in</strong>dω 1 = ω 2 ∧ ω 21 = −ω 2 ∧ ω 12 = ω 12 ∧ ω 2 <strong>and</strong> dω 2 = ω 1 ∧ ω 12 .( 3∑ )0 = d(de i )=d ω ik e k ==k=13∑dω ij e j −j=13∑k=1(dω ik e k − ω ik ∧3∑ ( 3∑ω ik ∧ ω kj)e j =j=1∑so dω ij − 3 ω ik ∧ ω kj =0for all i, j.k=1k=1We also have the follow<strong>in</strong>g additional consequence of the proof:3∑ )ω kj e jj=13∑j=1(dω ij −Proposition 3.2. The shape operator is symmetric, i.e., h 12 = h 21 .□3∑ω ik ∧ ω kj)e j ,Proof. From the e 3 component of the equation d(dx) =0 <strong>in</strong> the proof of Theorem 3.1 we have0=ω 1 ∧ ω 13 + ω 2 ∧ ω 23 = ω 1 ∧ (h 11 ω 1 + h 12 ω 2 )+ω 2 ∧ (h 21 ω 1 + h 22 ω 2 )=(h 12 − h 21 )ω 1 ∧ ω 2 ,k=1so h 12 − h 21 =0.□Recall that V is a pr<strong>in</strong>cipal direction if de 3 (V) isascalar multiple of V. So e 1 <strong>and</strong> e 2 arepr<strong>in</strong>cipal directions if <strong>and</strong> only if h 12 =0<strong>and</strong> we have ω 13 = k 1 ω 1 <strong>and</strong> ω 23 = k 2 ω 2 , where k 1 <strong>and</strong>k 2 are, as usual, the pr<strong>in</strong>cipal curvatures.It is important to underst<strong>and</strong> how our battery of forms changes if we change our mov<strong>in</strong>g frameby rotat<strong>in</strong>g e 1 , e 2 through some angle θ (which may be a function).Lemma 3.3. Suppose e 1 = cos θe 1 + s<strong>in</strong> θe 2 <strong>and</strong> e 2 = − s<strong>in</strong> θe 1 + cos θe 2 for some function θ.Then we haveω 1 = cos θω 1 + s<strong>in</strong> θω 2ω 2 = − s<strong>in</strong> θω 1 + cos θω 2ω 12 = ω 12 + dθω 13 = cos θω 13 + s<strong>in</strong> θω 23ω 23 = − s<strong>in</strong> θω 13 + cos θω 23Note, <strong>in</strong> particular, that ω 1 ∧ ω 2 = ω 1 ∧ ω 2 <strong>and</strong> ω 13 ∧ ω 23 = ω 13 ∧ ω 23 .Proof. We leave this to the reader <strong>in</strong> Exercise 1.□It is often convenient when we study curves <strong>in</strong> surfaces (as we did <strong>in</strong> Sections 3 <strong>and</strong> 4 of Chapter2) to use the Darboux frame, a mov<strong>in</strong>g frame for the surface adapted so that e 1 is tangent to thecurve. (See Exercise 3.) For example, α is a geodesic if <strong>and</strong> only if <strong>in</strong> terms of the Darboux framewe have ω 12 =0as a 1-form on α.Let’s now exam<strong>in</strong>e the structure equations more carefully.


98 Chapter 3. <strong>Surfaces</strong>: Further TopicsGauss equation: dω 12 = −ω 13 ∧ ω 23Codazzi equations: dω 13 = ω 12 ∧ ω 23dω 23 = −ω 12 ∧ ω 13Example 1. To illustrate the power of the mov<strong>in</strong>g frame approach, we reprove Proposition 3.3of Chapter 2: Suppose K =0<strong>and</strong> M has no planar po<strong>in</strong>ts. Then we claim that M is ruled <strong>and</strong> thetangent plane of M is constant along the rul<strong>in</strong>gs. We work <strong>in</strong> a pr<strong>in</strong>cipal mov<strong>in</strong>g frame with k 1 =0,so ω 13 =0. Therefore, by the first Codazzi equation, dω 13 =0=ω 12 ∧ ω 23 = ω 12 ∧ k 2 ω 2 . S<strong>in</strong>cek 2 ≠0,wemust have ω 12 ∧ ω 2 =0,<strong>and</strong> so ω 12 = fω 2 for some function f. Therefore, ω 12 (e 1 )=0,<strong>and</strong> so de 1 (e 1 )=ω 12 (e 1 )e 2 + ω 13 (e 1 )e 3 = 0. It follows that e 1 stays constant as we move <strong>in</strong> thee 1 direction, so follow<strong>in</strong>g the e 1 direction gives us a l<strong>in</strong>e. Moreover, de 3 (e 1 )=0 (s<strong>in</strong>ce k 1 = 0), sothe tangent plane to M is constant along that l<strong>in</strong>e. ▽The Gauss equation is particularly <strong>in</strong>terest<strong>in</strong>g. <strong>First</strong>, note thatω 13 ∧ ω 23 =(h 11 ω 1 + h 12 ω 2 ) ∧ (h 12 ω 1 + h 22 ω 2 )=(h 11 h 22 − h 2 12)ω 1 ∧ ω 2 = KdA,where K = det [ h αβ]= det SP is the Gaussian curvature. So, the Gauss equation really reads:(⋆)dω 12 = −KdA.(How elegant!) Note, moreover, that, by Lemma 3.3, for any two mov<strong>in</strong>g frames e 1 , e 2 , e 3 <strong>and</strong>e 1 , e 2 , e 3 ,wehave dω 12 = dω 12 (which is good, s<strong>in</strong>ce the right-h<strong>and</strong> side of (⋆) doesn’t depend onthe frame field). Next, we observe that, because of the first equations <strong>in</strong> Theorem 3.1, ω 12 canbe computed just from know<strong>in</strong>g ω 1 <strong>and</strong> ω 2 , hence depends just on the first fundamental form ofthe surface. (If we write ω 12 = Pω 1 + Qω 2 , then the first equation determ<strong>in</strong>es P <strong>and</strong> the seconddeterm<strong>in</strong>es Q.) We therefore arrive at a new proof of Gauss’s Theorema Egregium, Theorem 3.1of Chapter 2.Example 2. Let’s go back to our usual parametrization of the unit sphere,Then we havex(u, v) =(s<strong>in</strong> u cos v, s<strong>in</strong> u s<strong>in</strong> v, cos u), 0


§3. Surface Theory with Differential Forms 99The 1-form ω 12 is called the connection form <strong>and</strong> measures the tangential twist of e 1 . Just aswe saw <strong>in</strong> Section 1, then, ∇ V e 1 is the tangential component of D V e 1 = de 1 (V) =ω 12 (V)e 2 +ω 13 (V)e 3 , which is, of course, ω 12 (V)e 2 .From the Gauss equation <strong>and</strong> Stokes’s Theorem, the Gauss-Bonnet formula follows immediatelyfor an oriented surface M with (piecewise smooth) boundary ∂M on which we can globally def<strong>in</strong>ea mov<strong>in</strong>g frame. That is, we can reprove the Local Gauss-Bonnet formula, Theorem 1.5, quiteeffortlessly.Proof. We start with an arbitrary mov<strong>in</strong>g frame e 1 , e 2 , e 3 <strong>and</strong> take a Darboux frame e 1 , e 2 , e 3along ∂M. Wewrite e 1 = cos θe 1 +s<strong>in</strong> θe 2 <strong>and</strong> e 2 = − s<strong>in</strong> θe 1 +cos θe 2 (where θ is smoothly chosenalong the smooth pieces of ∂M <strong>and</strong> the exterior angle ɛ j at P j gives the “jump” of θ as we crossP j ). Then, by Stokes’s Theorem <strong>and</strong> Lemma 3.3, we have∫∫∫∫∫∫(KdA = − dω 12 = − ω 12 = − ω12 − dθ ) ∫= − κ g ds +(2π − ∑ ɛ j ).M(See Exercise 2.)□M∂M∂M∂MEXERCISES 3.31. Prove Lemma 3.3.2. Let e 1 , e 2 , e 3 be the Darboux frame along a curve α. Show that as a 1-form on α, ω 12 = κ g ω 1 .3. Suppose α is a curve ly<strong>in</strong>g <strong>in</strong> the surface M. Let e 1 , e 2 , e 3 be the Darboux frame along α (i.e.,amov<strong>in</strong>g frame for the surface with e 1 tangent to α), <strong>and</strong> let e 1 = e 1 , e 2 , e 3 be the Frenetframe. Then, by analogy with Lemma 3.3, e 2 , e 3 are obta<strong>in</strong>ed from e 2 , e 3 by rotat<strong>in</strong>g throughsome angle θ. Show that, as 1-forms on α, wehave:ω 12 = κω 1 = cos θω 12 + s<strong>in</strong> θω 13ω 13 = 0 = − s<strong>in</strong> θω 12 + cos θω 13ω 23 = τω 1 = ω 23 + dθ.*4. Use Exercise 3 to prove Meusnier’s Theorem (Proposition 2.5 of Chapter 2).5. Use Exercise 3 to prove that if C ⊂ M is a l<strong>in</strong>e of curvature <strong>and</strong> the osculat<strong>in</strong>g plane of Cmakes a constant angle with the tangent plane of M, then C is planar.6. Us<strong>in</strong>g mov<strong>in</strong>g frames to redo Exercise 2.2.13.*7. Use mov<strong>in</strong>g frames to compute the Gaussian curvature of the torus, parametrized as <strong>in</strong> Example1(c) of Chapter 2.8. The vectors e 1 = v(1, 0) <strong>and</strong> e 2 = v(0, 1) give a mov<strong>in</strong>g frame at (u, v) ∈ H. Set ω 1 = du/v<strong>and</strong> ω 2 = dv/v.a. Check that for any V ∈ T (u,v) H, ω 1 (V) =I(V, e 1 ) <strong>and</strong> ω 2 (V) =I(V, e 2 ).


100 Chapter 3. <strong>Surfaces</strong>: Further Topicsb. Compute ω 12 <strong>and</strong> dω 12 <strong>and</strong> verify that K = −1.9. Use mov<strong>in</strong>g frames to redoa. Exercise 3.1.9b. Exercise 3.1.1010. a. Use mov<strong>in</strong>g frames to reprove the result of Exercise 2.3.13.b. Suppose there are two families of geodesics <strong>in</strong> M mak<strong>in</strong>g a constant angle θ. Prove ordisprove: M is flat.11. Recall that locally any 1-form φ with dφ =0can be written <strong>in</strong> the form φ = df for somefunction f.a. Prove that if a surface M is flat, then locally we can f<strong>in</strong>d a mov<strong>in</strong>g frame e 1 , e 2 on M sothat ω 12 =0. (H<strong>in</strong>t: Start with an arbitrary mov<strong>in</strong>g frame.)b. Deduce that if M is flat, locally we can f<strong>in</strong>d a parametrization x of M with E = G =1<strong>and</strong> F =0. (That is, locally M is isometric to a plane.)12. (The Bäcklund transform) Suppose M <strong>and</strong> M are two surfaces <strong>in</strong> R 3 <strong>and</strong> f : M → M is asmooth bijective function with the properties that(i) the l<strong>in</strong>e from P to f(P )istangent to M at P <strong>and</strong> tangent to M at f(P );(ii) the distance between P <strong>and</strong> f(P )isaconstant r, <strong>in</strong>dependent of P ;(iii) the angle between n(P ) <strong>and</strong> n(f(P )) is a constant θ, <strong>in</strong>dependent of P .Prove that both M <strong>and</strong> M have constant curvature K = −(s<strong>in</strong> 2 θ)/r 2 . (H<strong>in</strong>ts: Write P = f(P ),<strong>and</strong> let e 1 , e 2 , e 3 (resp. e 1 , e 2 , e 3 )bemov<strong>in</strong>g frames at P (resp. P ) with e 1 = e 1 <strong>in</strong> the directionof −→P P . Lett<strong>in</strong>g x <strong>and</strong> x = f ◦x be local parametrizations, we have x = x + re 1 . Differentiatethis equation <strong>and</strong> deduce that ω 12 = cot θω 13 − 1 r ω 2.)4. Calculus of Variations <strong>and</strong> <strong>Surfaces</strong> of Constant Mean CurvatureEvery student of calculus is familiar with the necessary condition for a differentiable functionf : R n → R to have a local extreme po<strong>in</strong>t (m<strong>in</strong>imum or maximum) at P :Wemust have ∇f(P )=0.Phrased slightly differently, for every vector V, the directional derivativef(P + εV) − f(P )D V f(P )=limε→0 εshould vanish. Moreover, if we are given a constra<strong>in</strong>t set M = {x ∈ R n : g 1 (x) =0,g 2 (x) =0,...,g k (x) =0}, the method of Lagrange multipliers tells us that at a constra<strong>in</strong>ed extreme po<strong>in</strong>tP we must havek∑∇f(P )= λ i ∇g i (P )i=1for some scalars λ 1 ,...,λ k . (There is also a nondegeneracy hypothesis here that ∇g 1 (P ),...,∇g k (P )be l<strong>in</strong>early <strong>in</strong>dependent.)Suppose we are given a regular parametrized surface x: U → R 3 <strong>and</strong> want to f<strong>in</strong>d—without thebenefit of the analysis of Section 4 of Chapter 2—a geodesic from P = x(u 0 ,v 0 )toQ = x(u 1 ,v 1 ).


§4. Calculus of Variations <strong>and</strong> <strong>Surfaces</strong> of Constant Mean Curvature 101Among all paths α: [0, 1] → M with α(0) = P <strong>and</strong> α(1) = Q, wewish to f<strong>in</strong>d the shortest. Thatis, we want to choose the path α(t) =x(u(t),v(t)) so as to m<strong>in</strong>imize the <strong>in</strong>tegral∫ 10‖α ′ (t)‖dt =∫ 10√E(u(t),v(t))(u ′ (t)) 2 +2F (u(t),v(t))u ′ (t)v ′ (t)+G(u(t),v(t))(v ′ (t)) 2 dtsubject to the constra<strong>in</strong>ts that (u(0),v(0)) = (u 0 ,v 0 ) <strong>and</strong> (u(1),v(1)) = (u 1 ,v 1 ), as <strong>in</strong>dicated <strong>in</strong>Figure 4.1. Now we’re do<strong>in</strong>g a m<strong>in</strong>imization problem <strong>in</strong> the space of all (C 1 ) curves (u(t),v(t)) with(u 1 ,v 1 ) PQ(u 0 ,v 0 )Figure 4.1(u(0),v(0)) = (u 0 ,v 0 ) <strong>and</strong> (u(1),v(1)) = (u 1 ,v 1 ). Even though we’re now work<strong>in</strong>g <strong>in</strong> an <strong>in</strong>f<strong>in</strong>itedimensionalsett<strong>in</strong>g, we should not panic. In classical term<strong>in</strong>ology, we have a functional F def<strong>in</strong>edon the space X of C 1 curves u: [0, 1] → R 3 , i.e.,(∗)F (u) =∫ 10f(t, u(t), u ′ (t))dt.For example, <strong>in</strong> the case of the arclength problem, we havef ( t, (u(t),v(t)), (u ′ (t),v ′ (t)) ) =√E(u(t),v(t))(u ′ (t)) 2 +2F (u(t),v(t))u ′ (t)v ′ (t)+G(u(t),v(t))(v ′ (t)) 2 .To say that a particular curve u ∗ is a local extreme po<strong>in</strong>t (with fixed endpo<strong>in</strong>ts) of the functionalF given <strong>in</strong> (∗) istosay that for any variation ξ :[0, 1] → R 2 with ξ(0) = ξ(1) = 0, the directionalderivativeshould vanish. This leads us to theD ξ F (u ∗ )=limε→0F (u ∗ + εξ) − F (u ∗ )ε= d dε∣ F (u ∗ + εξ)ε=0Theorem 4.1 (Euler-Lagrange Equations). If u ∗ is a local extreme po<strong>in</strong>t of the functional Fgiven above <strong>in</strong> (∗), then at u ∗ we have∂f∂u = d ( ) ∂fdt ∂u ′ ,evaluat<strong>in</strong>g these both at (t, u ∗ (t), u ∗′ (t)), for all 0 ≤ t ≤ 1.


102 Chapter 3. <strong>Surfaces</strong>: Further TopicsProof. Let ξ :[0, 1] → R 2 be a C 1 curve with ξ(0) = ξ(1) = 0. Then, us<strong>in</strong>g the fact that wecan pull the derivative under the <strong>in</strong>tegral sign (see Exercise 1) <strong>and</strong> then the cha<strong>in</strong> rule, we haveddε∣ F (u ∗ + εξ) = d ∫ 1ε=0dε∣ f(t, u ∗ (t)+εξ(t), u ∗′ (t)+εξ ′ (t))dtε=0 0∫ 1∂=0 ∂ε∣ f(t, u ∗ (t)+εξ(t), u ∗′ (t)+εξ ′ (t))dtε=0∫ 1 ( ∂f=∂u (t, u∗ (t), u ∗′ (t)) · ξ(t)+ ∂f)∂u ′ (t, u∗ (t), u ∗′ (t)) · ξ ′ (t) dt<strong>and</strong> so, <strong>in</strong>tegrat<strong>in</strong>g by parts, we have==0∫ 10∫ 10( ∂f∂u · ξ(t) − d dt( ∂f∂u − d dt( ) ∂f)∂u ′ · ξ(t) dt + ∂f ] 1∂u ′ · ξ(t) 0( ∂f∂u ′ ) )· ξ(t)dt.Now, apply<strong>in</strong>g Exercise 2, s<strong>in</strong>ce this holds for all C 1 ξ with ξ(0) = ξ(1) = 0, we<strong>in</strong>fer that∂f∂u − d ( ) ∂fdt ∂u ′ = 0,as desired. □Of course, the Euler-Lagrange equations really give a system of differential equations:∂f∂u = d ( ) ∂fdt ∂u(♣)′ ∂f∂v = d ( ) ∂fdt ∂v ′ .Example 1. Recall that for the unit sphere <strong>in</strong> the usual parametrization we have E =1,F =0,<strong>and</strong> G = s<strong>in</strong> 2 u. To f<strong>in</strong>d the shortest path from (u 0 ,v 0 )=(u 0 ,v 0 )tothepo<strong>in</strong>t (u 1 ,v 1 )=(u 1 ,v 0 ),we want to m<strong>in</strong>imize the functionalF (u, v) =∫ 10√(u ′ (t)) 2 + s<strong>in</strong> 2 u(t)(v ′ (t)) 2 dt.Assum<strong>in</strong>g our critical path u ∗ is parametrized at constant speed, the equations (♣) give us v ′ (t) =const <strong>and</strong> u ′′ (t) =s<strong>in</strong> u(t) cos u(t)v ′ (t) 2 . (Cf. Example 6(b) <strong>in</strong> Section 4 of Chapter 2.) ▽We now come to two problems that <strong>in</strong>terest us here: What is the surface of least area with agiven boundary curve? And what is the surface of least area conta<strong>in</strong><strong>in</strong>g a given volume? For thiswe must consider parametrized surfaces <strong>and</strong> hence functionals def<strong>in</strong>ed on functions of two variables.In particular, for functions x: D → R 3 def<strong>in</strong>ed on a given doma<strong>in</strong> D ⊂ R 2 ,weconsider∫∫F (x) = ‖x u × x v ‖dudv.We seek a function x ∗ so that, for all variations ξ : D → R 3 with ξ = 0 on ∂D,D ξ F (u ∗ F (u ∗ + εξ) − F (u ∗ ))=lim= d ε→0 εdε∣ F (u ∗ + εξ) =0.ε=0D


§4. Calculus of Variations <strong>and</strong> <strong>Surfaces</strong> of Constant Mean Curvature 103Now we compute: Recall<strong>in</strong>g that d dt ‖f(t)‖ = f(t) · f ′ (t)<strong>and</strong> sett<strong>in</strong>g x = x ∗ + εξ, wehave‖f(t)‖d1 (dε∣ ‖x u × x v ‖ =(ξuε=0‖x ∗ u × x ∗ × x ∗ v + x ∗ u × ξ v ) · (x ∗ u × x ∗ v) )v‖Next we observe that=(ξ u × x ∗ v + x ∗ u × ξ v ) · n.(ξ u × x ∗ v) · n = ( (ξ × x ∗ v) · n ) u − (ξ × x∗ uv) · n − (ξ × x ∗ v) · n u(x ∗ u × ξ v ) · n = ( (x ∗ u × ξ) · n ) v − (x∗ uv × ξ) · n − (x ∗ u × ξ) · n v ,<strong>and</strong> so, add<strong>in</strong>g these equations, we obta<strong>in</strong>(ξ u × x ∗ v + x ∗ u × ξ v ) · n = ( (ξ × x ∗ v) · n ) u + ( (x ∗ u × ξ) · n ) v − ( (ξ × x ∗ v) · n u +(x ∗ u × ξ) · n v)= ( (ξ × x ∗ v) · n ) u − ( (ξ × x ∗ u) · n ) v − ( (ξ × x ∗ v) · n u +(x ∗ u × ξ) · n v)= ( (ξ × x ∗ v) · n ) u − ( (ξ × x ∗ u) · n ) v − ξ · (x ∗ v × n u + n v × x ∗ u).At the last step, we’ve used the identity (U × V) · W =(W × U) · V =(V × W) · U. Theappropriate way to <strong>in</strong>tegrate by parts <strong>in</strong> the two-dimensional sett<strong>in</strong>g is to apply Green’s Theorem,Theorem 2.6 of the Appendix, <strong>and</strong> so we let P =(ξ × x ∗ u) · n <strong>and</strong> Q =(ξ × x ∗ v) · n <strong>and</strong> obta<strong>in</strong>∫∫(ξ u × x ∗ v + x ∗ u × ξ v ) · ndudvD∫∫ ( ((ξ= × x∗v ) · n ) − ( (ξ × x ∗ uu) · n ) ) ∫∫dudv − ξ · (x ∗vv × n u + n v × x ∗ u)dudvD } {{ } } {{ }DQ uP v∫=∂D(ξ × x ∗ u) · n du +(ξ × x ∗} {{ }v) · n} {{ }PQ∫∫dv − ξ · (x ∗ v × n u + n v × x ∗ u)dudv.DS<strong>in</strong>ce ξ = 0 on ∂D, the l<strong>in</strong>e <strong>in</strong>tegral vanishes. Us<strong>in</strong>g the equations (††) onp.56, we f<strong>in</strong>d thatx ∗ v × n u = a(x ∗ u × x ∗ v) <strong>and</strong> n v × x ∗ u = d(x ∗ u × x ∗ v), so, at long last, we obta<strong>in</strong>∫∫∫∫ddε∣ ‖x u × x v ‖dudv = (ξ u × x ∗ v + x ∗ u × ξ v ) · ndudvε=0 DD∫∫= − ξ · (x ∗ v × n u + n v × x ∗ u)dudvD∫∫∫∫= − (a + d)ξ · (x ∗ u × x ∗ v)dudv = − 2Hξ · ndA,DDs<strong>in</strong>ce H = 1 2 trS P .We conclude from this, us<strong>in</strong>g a two-dimensional analogue of Exercise 2, the follow<strong>in</strong>gTheorem 4.2. Among all (parametrized) surfaces with a given boundary curve, the one ofleast area is m<strong>in</strong>imal, i.e., has H =0.This result, <strong>in</strong>deed, is the orig<strong>in</strong> of the term<strong>in</strong>ology.Next, suppose we wish to characterize those closed surfaces of least area conta<strong>in</strong><strong>in</strong>g a givenvolume V .Tomake a parametrized surface closed, we require that x(u, v) =x 0 for all (u, v) ∈ ∂D.But how do we express the volume constra<strong>in</strong>t <strong>in</strong> terms of x? The answer comes from the Divergence


104 Chapter 3. <strong>Surfaces</strong>: Further TopicsTheorem <strong>and</strong> is the three-dimensional analogue of the result of Exercise A.2.5: The volume enclosedby the parametrized surface x is given byvol(V )= 1 ∫∫x · ndA.3 DThus, the method of ∫∫Lagrange multipliers suggests that for a surface of least area there must be aconstant λ so that (2H − λ)ξ · ndA =0for all variations ξ with ξ = 0 on ∂D. Once aga<strong>in</strong>,Dus<strong>in</strong>g a two-dimensional analogue of Exercise 2, we see that 2H − λ =0<strong>and</strong> hence H must beconstant. (Also see Exercise 6.) We conclude:Theorem 4.3. Among all (parametrized) surfaces conta<strong>in</strong><strong>in</strong>g a fixed volume, the one of leastarea has constant mean curvature.In particular, a soap bubble should have constant mean curvature. A nontrivial theorem ofAlex<strong>and</strong>rov, analogous to Theorem 3.5 of Chapter 2, states that a smooth, compact surface ofconstant mean curvature must be a sphere. So soap bubbles should be spheres. How do youexpla<strong>in</strong> “double bubbles”?Example 2. If we ask which surfaces of revolution have constant mean curvature H 0 , thestatement of Exercise 2.2.19a. leads us to the differential equationh ′′(1 + h ′2 ) 3/2 − 1h(1 + h ′2 ) 1/2 =2H 0.(Here the surface is obta<strong>in</strong>ed by rotat<strong>in</strong>g the graph of h about the coord<strong>in</strong>ate axis.) We can rewritethis equation as follows:<strong>and</strong>, multiply<strong>in</strong>g through by h ′ ,−hh ′′ +(1+h ′2 )(1 + h ′2 ) 3/2 +2H 0 h =0(†)h ′ −hh′′ +(1+h ′2 )(1 + h ′2 ) 3/2 +2H 0 hh ′ =0( )h ′(1√ +2H 01+h ′2 2 h2) ′ =0h√ + H 0h 2 = const.1+h ′2We now show that such functions have a wonderful geometric characterization, as suggested<strong>in</strong> Figure 4.2. Start<strong>in</strong>g with an ellipse with semimajor axis a <strong>and</strong> semim<strong>in</strong>or axis b, weconsiderFigure 4.2


§4. Calculus of Variations <strong>and</strong> <strong>Surfaces</strong> of Constant Mean Curvature 105the locus of one focus as we roll the ellipse along the x-axis. By def<strong>in</strong>ition of an ellipse, we have‖ −−→ F 1 Q‖ + ‖ −−→ F 2 Q‖ =2a, <strong>and</strong> by Exercise 7, we have yy 2 = b 2 (see Figure 4.3). On the other h<strong>and</strong>, wededuce from Exercise 8 that −−→ F 1 Q is normal to the curve, <strong>and</strong> that, therefore, y = ‖ −−→ F 1 Q‖ cos φ. S<strong>in</strong>cethe “reflectivity” property of the ellipse tells us that ∠F 1 QP 1∼ = ∠F2 QP 2 ,wehave y 2 = ‖ −−→ F 2 Q‖ cos φ.S<strong>in</strong>ce cos φ = dx/ds <strong>and</strong> ds/dx = √ 1+(dy/dx) 2 ,wehaveF 1φxyP 1F 2y 2Q P 2Figure 4.3y + b2y = y + y 2 =2a cos φ =2a dxds<strong>and</strong> so0=y 2 − 2ay dxds + b2 = y 2 − √ 2ay + 1+y ′2 b2 =0.Sett<strong>in</strong>g H 0 = −1/2a, wesee that this matches the equation (†) above.▽EXERCISES 3.4♯ 1. Suppose g :[0, 1] × (−1, 1) → R is cont<strong>in</strong>uous <strong>and</strong> let G(ε) = g(t, ε)dt. Prove that if ∂g0∂ε is∫ 1∫cont<strong>in</strong>uous, then G ′ ∂gε ∫ 1(0) =∂ε (t, 0)dt. (H<strong>in</strong>t: Consider h(ε) = ∂g(t, u)dtdu.)∂ε♯ 2. *a. Suppose f is a cont<strong>in</strong>uous function on [0, 1] <strong>and</strong>0∫ 1functions ξ on [0, 1]. Prove that f =0. (H<strong>in</strong>t: Take ξ = f.)b. Suppose f is a cont<strong>in</strong>uous function on [0, 1] <strong>and</strong>0∫ 10∫ 100f(t)ξ(t)dt = 0 for all cont<strong>in</strong>uousf(t)ξ(t)dt = 0 for all cont<strong>in</strong>uousfunctions ξ on [0, 1] with ξ(0) = ξ(1) = 0. Prove that f =0. (H<strong>in</strong>t: Take ξ = ψf for anappropriate cont<strong>in</strong>uous function ψ.)c. Deduce the same result for C 1 functions ξ.d. Deduce the same result for vector-valued functions f <strong>and</strong> ξ.3. Use the Euler-Lagrange equations to show that the shortest path jo<strong>in</strong><strong>in</strong>g two po<strong>in</strong>ts <strong>in</strong> theEuclidean plane is a l<strong>in</strong>e segment.


106 Chapter 3. <strong>Surfaces</strong>: Further Topics4. Use the functional F (u) =∫ ba2πu(t) √ 1+(u ′ (t)) 2 dt to determ<strong>in</strong>e the surface of revolution ofleast area with two parallel circles (perhaps of different radii) as boundary. (H<strong>in</strong>t: You shouldend up with the same differential equation as <strong>in</strong> Exercise 2.2.19.)5. Prove the analogue of Theorem 4.3 for curves. That is, show that of all closed plane curvesenclos<strong>in</strong>g a given area, the circle has the least perimeter. (Cf. Theorem 3.10 of Chapter 1. H<strong>in</strong>t:Start with Exercise A.2.5. Show that the constra<strong>in</strong>ed Euler-Lagrange equations imply that theextremiz<strong>in</strong>g curve has constant curvature. Exercise 1.2.4 <strong>and</strong> the cha<strong>in</strong> rule will help.)6. Interpret<strong>in</strong>g the <strong>in</strong>tegral∫ 10space of cont<strong>in</strong>uous functions on [0, 1], prove that iffunctions g with〈f ⊥ , 1〉 = 0.)∫ 10f(t)g(t)dt as an <strong>in</strong>ner product (dot product) 〈f,g〉 on the vector∫ 10f(t)g(t)dt = 0 for all cont<strong>in</strong>uousg(t)dt =0,then f must be constant. (H<strong>in</strong>t: Write f = 〈f,1〉1+f ⊥ , where7. Prove the pedal property of the ellipse: The product of the distances from the foci to the tangentl<strong>in</strong>e of the ellipse at any po<strong>in</strong>t is a constant (<strong>in</strong> fact, the square of the semim<strong>in</strong>or axis).8. The curve α(s) rolls along the x-axis, start<strong>in</strong>g at the po<strong>in</strong>t α(0) = 0. Apo<strong>in</strong>t F is fixed relativeto the curve. Let β(s) bethe curve that F traces out. As <strong>in</strong>dicated <strong>in</strong> Figure 4.4, let θ(s) beFθQF'Q'Figure 4.4[]the angle α ′ cos θ − s<strong>in</strong> θ(s) makes with the positive x-axis. Denote by R θ =the matrixs<strong>in</strong> θ cos θthat gives rotation of the plane through angle θ.a. Show that β(s) =(s, 0) + R −θ(s) (F − α(s)).b. Show that β ′ (s) · R −θ(s) (F − α(s)) = 0. That is, as F moves, <strong>in</strong>stantaneously it rotatesabout the contact po<strong>in</strong>t on the x-axis. (Cf. Exercise A.1.4.)9. F<strong>in</strong>d the path followed by the focus of the parabola y = x 2 /2asthe parabola rolls along thex-axis. The focus is orig<strong>in</strong>ally at (0, 1/2). (H<strong>in</strong>t: See Example 2.)


APPENDIXReview of L<strong>in</strong>ear Algebra <strong>and</strong> Calculus1. L<strong>in</strong>ear Algebra ReviewRecall that the set {v 1 ,...,v k } of vectors <strong>in</strong> R n gives a basis for a subspace V of R n if <strong>and</strong>only if every vector v ∈ V can be written uniquely as a l<strong>in</strong>ear comb<strong>in</strong>ation v = c 1 v 1 + ···+ c k v k .In particular, v 1 ,...,v n will form a basis for R n if <strong>and</strong> only if the n × n matrix⎡⎤| | |⎢⎥A = ⎣ v 1 v 2 ··· v n ⎦| | |is <strong>in</strong>vertible, <strong>and</strong> are said to be positively oriented if the determ<strong>in</strong>ant det A is positive. In particular,given two l<strong>in</strong>early <strong>in</strong>dependent vectors v, w ∈ R 3 , the set {v, w, v × w} always gives a positivelyoriented basis for R 3 .We say e 1 ,...,e k ∈ R n form an orthonormal set <strong>in</strong> R n if e i · e j =0for all i ≠ j <strong>and</strong> ‖e i ‖ =1for all i =1,...,k. Then we have the follow<strong>in</strong>gProposition 1.1. If {e 1 ,...,e n } is an orthonormal set of vectors <strong>in</strong> R n , then they form a basisfor R n <strong>and</strong>, given any v ∈ R n ∑,wehavev = n (v · e i )e i .i=1We sayann × n matrix A is orthogonal if A T A = I. Itiseasy to check that the column vectorsof A form an orthonormal basis for R n (<strong>and</strong> the same for the row vectors). Moreover, from thebasic formula Ax·y = x·A T y we deduce that if e 1 ,...,e k form an orthonormal set of vectors <strong>in</strong> R n<strong>and</strong> A is an orthogonal n × n matrix, then Ae 1 ,...,Ae k are likewise an orthonormal set of vectors.An important issue for differential geometry is to identify the isometries of R 3 (although thesame argument will work <strong>in</strong> any dimension). Recall that an isometry of R 3 is a function f : R 3 → R 3so that for any x, y ∈ R 3 ,wehave‖f(x) − f(y)‖ = ‖x − y‖. Wenowprove theTheorem 1.2. Any isometry f of R 3 can be written <strong>in</strong> the form f(x) =Ax + c for someorthogonal 3 × 3 matrix A <strong>and</strong> some vector c ∈ R 3 .Proof. Let f(0) =c, <strong>and</strong> replace f with the function f − c. Ittoo is an isometry (why?) <strong>and</strong>fixes the orig<strong>in</strong>. Then ‖f(x)‖ = ‖f(x) − f(0)‖ = ‖x − 0‖ = ‖x‖, sothat f preserves lengths ofvectors. Us<strong>in</strong>g this fact, we prove that f(x) · f(y) =x · y for all x, y ∈ R 3 .Wehave‖f(x) − f(y)‖ 2 = ‖x − y‖ 2 =(x − y) · (x − y) =‖x‖ 2 − 2x · y + ‖y‖ 2 ;on the other h<strong>and</strong>, <strong>in</strong> a similar fashion,‖f(x) − f(y)‖ 2 = ‖f(x)‖ 2 − 2f(x) · f(y)+‖f(y)‖ 2 = ‖x‖ 2 − 2f(x) · f(y)+‖y‖ 2 .107


108 Appendix. Review of L<strong>in</strong>ear Algebra <strong>and</strong> CalculusWe conclude that f(x) · f(y) =x · y, asdesired.We next prove that f must be a l<strong>in</strong>ear function. Let {e 1 , e 2 , e 3 } be the st<strong>and</strong>ard orthonormalbasis for R 3 , <strong>and</strong> let f(e j ) = v j , j = 1, 2, 3. It follows from what we’ve already proved that{v 1 , v 2 , v 3 } is also an orthonormal basis. Given an arbitrary vector x ∈ R 3 ∑, write x = 3 x i e i <strong>and</strong>f(x) =3 ∑j=1y j v j . Then it follows from Proposition 1.1 thaty i = f(x) · v i = x · e i = x i ,so f is <strong>in</strong> fact l<strong>in</strong>ear. The matrix A represent<strong>in</strong>g f with respect to the st<strong>and</strong>ard basis has as its j thcolumn the vector v j . Therefore, by our earlier remarks, A is an orthogonal matrix, as required. □Indeed, recall that if T : R n → R n is a l<strong>in</strong>ear map <strong>and</strong> B = {v 1 ,...,v k } is a basis for R n ,then the matrix for T with respect to the basis B is the matrix whose j th column consists of thecoefficients of T (v j ) with respect to the basis B. That is, it is the matrixi=1A = [ a ij], where T (vj )=n∑a ij v i .i=1Recall that if A is an n×n matrix (or T : R n → R n is a l<strong>in</strong>ear map), a nonzero vector x is calledan eigenvector if Ax = λx (T (x) =λx, resp.) for some scalar λ, called the associated eigenvalue.[ ] a bTheorem 1.3. A symmetric 2 × 2 matrix A = (or symmetric l<strong>in</strong>ear map T : R 2 → R 2 )b chas two real eigenvalues λ 1 <strong>and</strong> λ 2 , <strong>and</strong>, provided λ 1 ≠ λ 2 , the correspond<strong>in</strong>g eigenvectors v 1 <strong>and</strong>v 2 are orthogonal.Proof. Consider the functionf : R 2 → R, f(x) =Ax · x = ax 2 1 +2bx 1 x 2 + cx 2 2.By the maximum value theorem, f has a m<strong>in</strong>imum <strong>and</strong> a maximum subject to the constra<strong>in</strong>tg(x) =x 2 1 + x2 2 =1. Say these occur, respectively, at v 1 <strong>and</strong> v 2 . By the method of Lagrangemultipliers, we <strong>in</strong>fer that there are scalars λ i so that ∇f(v i )=λ i ∇g(v i ), i =1, 2. By Exercise5, this means Av i = λ i v i , <strong>and</strong> so the Lagrange multipliers are actually the associated eigenvalues.Nowλ 1 (v 1 · v 2 )=Av 1 · v 2 = v 1 · Av 2 = λ 2 (v 1 · v 2 ).It follows that if λ 1 ≠ λ 2 ,wemust have v 1 · v 2 =0,asdesired.□We recall that, <strong>in</strong> practice, we f<strong>in</strong>d the eigenvalues by solv<strong>in</strong>g for the roots[of the]characteristica bpolynomial p(t) =det(A − tI). In the case of a symmetric 2 × 2 matrix A = ,weobta<strong>in</strong> theb cpolynomial p(t) =t 2 − (a + c)t +(ac − b 2 ), whose roots areλ 1 = 1 ( √(a + c) − (a − c)22 +4b 2) <strong>and</strong> λ 2 = 1 2((a + c)+√(a − c) 2 +4b 2) .


§2. Calculus Review 109EXERCISES A.1♯ *1.Suppose {v 1 , v 2 } gives a basis for R 2 . Given vectors x, y ∈ R 2 , prove that x = y if <strong>and</strong> only ifx · v i = y · v i , i =1, 2.*2. The geometric-arithmetic mean <strong>in</strong>equality states that√ab ≤a + b2for positive numbers a <strong>and</strong> b,with equality hold<strong>in</strong>g if <strong>and</strong> only if a = b. Give a one-l<strong>in</strong>e proof us<strong>in</strong>g the Cauchy-Schwarz<strong>in</strong>equality:|u · v| ≤‖u‖‖v‖ for vectors u <strong>and</strong> v ∈ R n ,with equality hold<strong>in</strong>g if <strong>and</strong> only if one is a scalar multiple of the other.3. Let w, x, y, z ∈ R 3 . Prove that(w × x) · (y × z) =(w · y)(x · z) − (w · z)(x · y).(H<strong>in</strong>t: Both sides are l<strong>in</strong>ear <strong>in</strong> each of the four variables, so it suffices to check the result onbasis vectors.)♯ 4. Suppose A(t) isadifferentiable family of 3 × 3 orthogonal matrices. Prove that A(t) −1 A ′ (t) isalways skew-symmetric.[ ] a b5. If A = is a symmetric 2 × 2 matrix, set f(x) =Ax · x <strong>and</strong> check that ∇f(x) =2Ax.b c2. Calculus ReviewRecall that a function f : U → R def<strong>in</strong>ed on an open subset U ⊂ R n is C k (k =0, 1, 2,...,∞)if all its partial derivatives of order ≤ k exist <strong>and</strong> are cont<strong>in</strong>uous on U. We will use the notation∂f∂u <strong>and</strong> f ∂ 2 fu <strong>in</strong>terchangeably, <strong>and</strong> similarly with higher order derivatives:∂v∂u = ∂ ( ∂f)is the∂v ∂usame as f uv , <strong>and</strong> so on.One of the extremely important results for differential geometry is the follow<strong>in</strong>gTheorem 2.1. If f is a C 2 function, then∂2 f∂u∂v = ∂2 f∂v∂u (or f uv = f vu ).The same results apply to vector-valued functions, work<strong>in</strong>g with component functions separately.If f : U → R is C 1 we can form its gradient by tak<strong>in</strong>g the vector ∇f = ( )f x1 ,f x2 ,...,f xn of itspartial derivatives. One of the most fundamental formulas <strong>in</strong> differential calculus is the cha<strong>in</strong> rule:Theorem 2.2. Suppose f : R n → R <strong>and</strong> α: R → R n are differentiable. Then (f ◦α) ′ (t) =∇f(α(t)) · α ′ (t).


110 Appendix. Review of L<strong>in</strong>ear Algebra <strong>and</strong> CalculusIn particular, if α(0) = P <strong>and</strong> α ′ (0) = V ∈ R n , then (f ◦α) ′ (0) = ∇f(P ) · V. This is somewhatsurpris<strong>in</strong>g, as the rate of change of f along α at P depends only on the tangent vector <strong>and</strong> onnoth<strong>in</strong>g more subtle about the curve.Proposition 2.3. D V f(P )=∇f(P ) · V. Thus, the directional derivative is a l<strong>in</strong>ear functionof V.Proof. If we take α(t) =P + tV, then by def<strong>in</strong>ition of the directional derivative, D V f(P )=(f ◦α) ′ (0) = ∇f(P ) · V. □Another important consequence of the cha<strong>in</strong> rule, essential throughout differential geometry, is thefollow<strong>in</strong>gProposition 2.4. Suppose S ⊂ R n is a subset with the property that any pair of po<strong>in</strong>ts of Scan be jo<strong>in</strong>ed by a C 1 curve. Then a C 1 function f : S → R with ∇f = 0 everywhere is a constantfunction.Proof. Fix P ∈ S <strong>and</strong> let Q ∈ S be arbitrary. Choose a C 1 curve α with α(0) = P <strong>and</strong>α(1) = Q. Then (f ◦α) ′ (t) =∇f(α(t)) · α ′ (t) =0for all t. Itisaconsequence of the Mean ValueTheorem <strong>in</strong> <strong>in</strong>troductory calculus that a function g :[0, 1] → R that is cont<strong>in</strong>uous on [0, 1] <strong>and</strong>has zero derivative throughout the <strong>in</strong>terval must be a constant. Therefore, f(Q) =(f ◦α)(1) =(f ◦α)(0) = f(P ). It follows that f must be constant on S. □We will also have plenty of occasion to use the vector version of the product rule:Proposition 2.5. Suppose f, g : R → R 3 are differentiable. Then we have(f · g) ′ (t) =f ′ (t) · g(t)+f(t) · g ′ (t) <strong>and</strong>(f × g) ′ (t) =f ′ (t) × g(t)+f(t) × g ′ (t).Last, from vector <strong>in</strong>tegral calculus, we recall the analogue of the Fundamental Theorem ofCalculus <strong>in</strong> R 2 :Theorem 2.6 (Green’s Theorem). Let R ⊂ R 2 be a region, <strong>and</strong> let ∂R denote its boundarycurve, oriented counterclockwise (i.e., so that the region is to its “left”). Then∫∫∫ ( ∂QP (u, v)du + Q(u, v)dv =∂u − ∂P )dudv.∂v∂RRdcR∂RabFigure 2.1


§2. Calculus Review 111Proof. We give the proof here just for the case where R is a rectangle. Take R =[a, b] × [c, d],as shown <strong>in</strong> Figure 2.1. Now we merely calculate, us<strong>in</strong>g the Fundamental Theorem of Calculusappropriately:∫∫R( ∂Q∂u − ∂P )dudv =∂v=∫ dc∫ dc∫ b=∫=a∂R(∫ ba)∂Q∂u du dv −∫ ba(∫ d(Q(b, v) − Q(a, v))dv −∫ bP (u, c)du +∫ dccaQ(b, v)dv −P (u, v)du + Q(u, v)dv,)∂P∂v dv du( )P (u, d) − P (u, c) du∫ baP (u, d)du −∫ dcQ(a, v)dvas required.□EXERCISES A.2♯ 1.Suppose f :(a, b) → R n is C 1 <strong>and</strong> nowhere zero. Prove that f/‖f‖ is constant if <strong>and</strong> only iff ′ (t) =λ(t)f(t) for some cont<strong>in</strong>uous scalar function λ. (H<strong>in</strong>t: Set g = f/‖f‖ <strong>and</strong> differentiate.Why must g ′ · g = 0?)2. Suppose α: (a, b) → R 3 is twice-differentiable <strong>and</strong> λ is a nowhere-zero twice differentiablescalar function. Prove that α, α ′ , <strong>and</strong> α ′′ are everywhere l<strong>in</strong>early <strong>in</strong>dependent if <strong>and</strong> only ifλα, (λα) ′ , <strong>and</strong> (λα) ′′ are everywhere l<strong>in</strong>early <strong>in</strong>dependent.3. Let f, g : R → R 3 be C 1 vector functions. Supposef ′ (t) =a(t)f(t)+b(t)g(t)g ′ (t) =c(t)f(t) − a(t)g(t)for some cont<strong>in</strong>uous functions a, b, <strong>and</strong> c. Prove that the parallelogram spanned by f(t) <strong>and</strong>g(t) lies <strong>in</strong> a fixed plane <strong>and</strong> has constant area.♯ *4.Prove that for any cont<strong>in</strong>uous vector-valued function f :[a, b] → R 3 ,wehave∥∫ ba∫ bf(t)dt∥ ≤ ‖f(t)‖dt.a♯ 5.Let R ⊂ R 2 be a region. Prove that∫area(R) =∂R∫udv = − vdu = 1 ∫−vdu + udv.∂R 2 ∂R


112 Appendix. Review of L<strong>in</strong>ear Algebra <strong>and</strong> Calculus3. Differential EquationsTheorem 3.1 (Fundamental Theorem of ODE’s). Suppose U ⊂ R n is open <strong>and</strong> I ⊂ R is anopen <strong>in</strong>terval conta<strong>in</strong><strong>in</strong>g 0. Suppose x 0 ∈ U. Iff : U × I → R n is cont<strong>in</strong>uous, then the differentialequationdxdt = f(x,t), x(0) = x 0has a unique solution x = x(t, x 0 ) def<strong>in</strong>ed for all t <strong>in</strong> some <strong>in</strong>terval I ′ ⊂ I. Moreover, If f is C k ,then x is C k as a function of both t <strong>and</strong> the <strong>in</strong>itial condition x 0 (def<strong>in</strong>ed for t <strong>in</strong> some <strong>in</strong>terval <strong>and</strong>x 0 <strong>in</strong> some open set).Of special <strong>in</strong>terest to us will be l<strong>in</strong>ear differential equations.Theorem 3.2. Suppose A(t) is a cont<strong>in</strong>uous n × n matrix function on an <strong>in</strong>terval I. Then thedifferential equationdxdt = A(t)x(t), x 0 = x 0 ,has a unique solution on the entire orig<strong>in</strong>al <strong>in</strong>terval I.For proofs of these, <strong>and</strong> related, theorems <strong>in</strong> differential equations, we refer the reader to anyst<strong>and</strong>ard differential equations text (e.g., Edwards <strong>and</strong> Penney, Boyce <strong>and</strong> DePrima, or Birkhoff<strong>and</strong> Rota).Theorem 3.3. Given two C k vector fields X <strong>and</strong> Y that are l<strong>in</strong>early <strong>in</strong>dependent on a neighborhoodU of 0 ∈ R 2 , locally we can choose coord<strong>in</strong>ates (u, v) on U ′ ⊂ U so that X is tangent to theu-curves (i.e., the curves v = const) <strong>and</strong> Y is tangent to the v-curves (i.e., the curves u = const).Proof. We make a l<strong>in</strong>ear change of coord<strong>in</strong>ates so that X(0) <strong>and</strong> Y(0) are the unit st<strong>and</strong>ardbasis vectors. Let x(t, x 0 )bethe solution of the differential equation dx/dt = X, x(0) = x 0 , givenby Theorem 3.1. On a neighborhood of 0, each po<strong>in</strong>t (x, y) can be written as(x, y) =x(t, (0,v))for some unique t <strong>and</strong> v, asillustrated <strong>in</strong> Figure 3.1. If we def<strong>in</strong>e the function f(t, v) =x(t, (0,v)) =coord<strong>in</strong>ates (u,v)(0,v)Y(0)y(s,(u,0))x(t,(0,v))X(0)(u,0)Figure 3.1(x(t, v),y(t, v)), we note that f t = X(f(t, v)) <strong>and</strong> f v (0, 0) =(0, 1), so the derivative matrix Df(0, 0)


§3. Differential Equations 113is the identity matrix. It follows from the Inverse Function Theorem that (locally) we can solve for(t, v) asaC k function of (x, y). Note that the level curves of v have tangent vector X, asdesired.Now we repeat this procedure with the vector field Y. Let y(s, y 0 )bethe solution of thedifferential equation dy/ds = Y <strong>and</strong> write(x, y) =y(s, (u, 0))for some unique s <strong>and</strong> u. We similarly obta<strong>in</strong> (s, u) locally as a C k function of (x, y). We claimthat (u, v) give the desired coord<strong>in</strong>ates. We only need to check that on a suitable neighborhoodof the orig<strong>in</strong> they are <strong>in</strong>dependent; but from our earlier discussion we have v x =0,v y =1attheorig<strong>in</strong>, <strong>and</strong>, analogously, u x =1<strong>and</strong> u y =0,aswell. Thus, the derivative matrix of (u, v) istheidentity at the orig<strong>in</strong> <strong>and</strong> the functions therefore give a local parametrization. □EXERCISES A.31. Suppose M(s) isadifferentiable 3 × 3 matrix function of s, K(s) isaskew-symmetric 3 × 3matrix function of s, <strong>and</strong>M ′ (s) =M(s)K(s), M(0) = O.Show that M(s) =O for all s by show<strong>in</strong>g that the trace of (M T M) ′ (s) isidentically 0.2. (Gronwall <strong>in</strong>equality <strong>and</strong> consequences)a. Suppose f :[a, b) → R is differentiable, nonnegative, <strong>and</strong> f(a) =c>0. Suppose g :[a, b) →R is cont<strong>in</strong>uous <strong>and</strong> f ′ (t) ≤ g(t)f(t) for all t. Prove that(∫ t)f(t) ≤ c exp g(u)du for all t.ab. Conclude that if f(a) =0,then f(t) =0for all t.c. Suppose now v :[a, b) → R n is a differentiable vector function, <strong>and</strong> M(t) isacont<strong>in</strong>uousn × n matrix function for t ∈ [a, b), <strong>and</strong> v ′ (t) =M(t)v(t). Apply the result of part bto conclude that if v(a) =0, then v(t) =0 for all t. Deduce uniqueness of solutions tol<strong>in</strong>ear first order differential equations for vector functions. (H<strong>in</strong>t: Let f(t) =‖v(t)‖ 2 <strong>and</strong>g(t) =2n max{|m ij (t)|}.)d. Use part c to deduce uniqueness of solutions to l<strong>in</strong>ear n th order differential equations.(H<strong>in</strong>t: Introduce new variables correspond<strong>in</strong>g to higher derivatives.)


ANSWERS TO SELECTED EXERCISES1.1.1 α(t) = ( 1−t 21+t 2 ,2t1+t 2 ).1.1.4 We parametrize the curve by α(t) = (t, f(t)), a ≤ t ≤ b, <strong>and</strong> so length(α) =∫ ba ‖α′ (t)‖ dt = ∫ ba√1+(f ′ (t)) 2 dt.1.1.6 β(s) = ( 12 (√ s 2 +4+s), 1 2 (√ s 2 +4− s), √ 2 ln(( √ s 2 +4+s)/2) ) .1.2.1 c. κ =12 √ 2 √ 1−s 21.2.3 a. T = 1 2 (√ 1+s, − √ 1 − s, √ 12), κ =2 √ 2 √ , N =1/√ 2( √ 1 − s, √ 1+s, 0), B =1−s 212 (−√ 1+s, √ 1 − s, √ 12), τ =2 √ 2 √ ; c. T = √ √1(t, √1−s 2 2 1+t1+t 2 , 1), κ = τ =21/2(1 + t 2 ), N = √ 1 (1, 0, −t), B = √ √1(−t, √1+t 2 2 1+t1+t 2 , −1)21.2.5 κ =1/ s<strong>in</strong>h t (which we see, once aga<strong>in</strong>, is the absolute value of the slope).1.2.6 B ′ =(T×N) ′ = T ′ ×N+T×N ′ =(κN)×N+T×(−κT+τB) =τ(T×B) =τ(−N),as required.1.2.9 b. If all the osculat<strong>in</strong>g planes pass through the orig<strong>in</strong>, then there are scalar functionsλ <strong>and</strong> µ so that 0 = α + λT + µN for all s. Differentiat<strong>in</strong>g <strong>and</strong> us<strong>in</strong>g the Frenetformulas, we obta<strong>in</strong> 0 = T + κλN + λ ′ T + µ ( −κT + τB ) + µ ′ N; collect<strong>in</strong>g terms, wehave 0 = ( 1+λ ′ − κµ ) T + ( κλ + µ ′) N + µτB for all s. S<strong>in</strong>ce {T, N, B} is a basisfor R 3 ,we<strong>in</strong>fer, <strong>in</strong> particular, that µτ =0. (We could also just have taken the dotproduct of the entire expression with B.) µ(s) =0leads to a contradiction, so wemust have τ =0<strong>and</strong> so the curve is planar.1.2.11 We have α ′ × α ′′ = κυ 3 B, so α ′ × α ′′′ = (α ′ × α ′′ ) ′ = (κυ 3 B) ′ = (κυ 3 ) ′ B +(κυ 3 )(−τυN), so (α ′ × α ′′′ ) · α ′′ = −κ 2 τυ 6 . Therefore, τ = α ′ · (α ′′ × α ′′′ )/(κ 2 υ 6 ),<strong>and</strong> <strong>in</strong>sert<strong>in</strong>g the formula of Proposition 2.2 gives the result.1.2.23 a. Consider the unit normal A s,t to the plane through P = 0, Q = α(s), <strong>and</strong>R = α(t). Choos<strong>in</strong>g coord<strong>in</strong>ates so that T(0) = (1, 0, 0), N(0) = (0, 1, 0), <strong>and</strong>B(0) = (0, 0, 1), we apply Proposition 2.6 to obta<strong>in</strong>α(s) × α(t) =st(s − t) (−κ212 0 τ 0 st + ...,2κ 0 τ 0 (s + t)+...,−6κ 0 +2κ ′ 0 (s + t) − κ3 0 st + ...) ,α(s) × α(t)so A s,t =→ A =(0, 0, −1) as s, t → 0. Thus, the plane through P‖α(s) × α(t)‖with normal A is the osculat<strong>in</strong>g plane.114


SELECTED ANSWERS 1151.2.23 a. cont. Alternatively, let the equation of the plane through P , Q, <strong>and</strong> R beA s,t · x =0(where we choose A s,t to vary cont<strong>in</strong>uously with length 1). We wantto determ<strong>in</strong>e A = lim s,t→0 A s,t . For fixed s <strong>and</strong> t, consider the function F s,t (u) =A s,t · α(u). Then F s,t (0) = F s,t (s) =F s,t (t) =0, so, by the mean value theorem,there are ξ 1 <strong>and</strong> ξ 2 so that F s,t(ξ ′ 1 )=F s,t(ξ ′ 2 )=0,hence η so that F s,t(η) ′′ =0. NowF s,t(0) ′ = A s,t · T(0) <strong>and</strong> F s,t(0) ′′ = A s,t · κ 0 N(0). S<strong>in</strong>ce ξ i → 0 <strong>and</strong> η → 0ass, t → 0,we obta<strong>in</strong> A · T(0) = A · N(0) = 0, so A = ±B(0), as desired.1.3.4 Let L = length(C). Then by Theorem 3.5 we have 2π = ∫ L0 κ(s) ds ≤ ∫ L0cds = cL,so L ≥ 2π/c.2.1.2 a. E = a 2 , F =0,G = a 2 s<strong>in</strong> 2 u; d. E = G = a 2 cosh 2 u, F =02.1.4 Say all the normal l<strong>in</strong>es pass through the orig<strong>in</strong>. Then there is a function λ so thatx = λn. Differentiat<strong>in</strong>g, we have x u = λn u + λ u n <strong>and</strong> x v = λn v + λ v n. Dott<strong>in</strong>g withn, weget 0 = λ u = λ v . Therefore, λ is a constant <strong>and</strong> so ‖x‖ = const. Alternatively,from the statement x = λn we proceed as follows. S<strong>in</strong>ce n · x u = n · x v =0,wehavex · x u = x · x v =0. Therefore, (x · x) u =(x · x) v =0,so‖x‖ 2 is constant.2.1.6 We check that E = G =4/(1 + u 2 + v 2 ) 2 <strong>and</strong> F =0,sothe result follows fromExercise 5.2.1.7 b. One of these is: x(u, v) =(cos u + v s<strong>in</strong> u, s<strong>in</strong> u − v cos u, v).2.1.12 a. If a cosh(1/a) =R, the area is 2π ( a + R √ R 2 − a 2) .2.2.1 If u- <strong>and</strong> v-curves are l<strong>in</strong>es of curvature, then F =0(because pr<strong>in</strong>cipal directionsare orthogonal) <strong>and</strong> m = S(x u ) · x v = k 1 x u · x v =0. Conversely, sett<strong>in</strong>g S P (x u )=ax u + bx v ,we<strong>in</strong>fer that if F = m =0,then 0 = S P (x u ) · x v = Fa+ Gb = Gb, <strong>and</strong>so b =0. Therefore, x u (<strong>and</strong>, similarly, x v )isaneigenvector for S P . Moreover, ifS P (x u )=k 1 x u <strong>and</strong> S P (x v )=k 2 x v ,wedot with x u <strong>and</strong> x v , respectively, to obta<strong>in</strong>l = Ek 1 <strong>and</strong> n = Gk 2 .[]1/b 02.2.3 b. l = b, m = 0, n = cos u(a + b cos u), S P =,0 cos u/(a + b cos u)H = 1 12(b + cos u )a+b cos u , K =cos ub(a+b cos u) ; d. l = −a, m = 0, n = a, S P =[]−(1/a)sech 2 u 00 (1/a)sech 2 , H =0,K = −(1/a) 2 sech 4 u.u2.2.5 We know from Example 1 of Chapter 1, Section 2 that the pr<strong>in</strong>cipal normal of thehelix po<strong>in</strong>ts along the rul<strong>in</strong>g <strong>and</strong> is therefore orthogonal to n. Aswemove along arul<strong>in</strong>g, n twists <strong>in</strong> a plane orthogonal to the rul<strong>in</strong>g, so its directional derivative <strong>in</strong> thedirection of the rul<strong>in</strong>g is orthogonal to the rul<strong>in</strong>g.2.2.6 E = tanh 2 u, F =0,G = sech 2 u, −l = sech u tanh u = n, m =02.3.5 d. Γ v uv =Γ vvu = f ′ (u)/f(u), Γ uvv = −f(u)f ′ (u), all others 0.2.4.4 κ g = cot u 0 ;wecan also deduce this from Figure 3.1, as the curvature vector κN =(1/ s<strong>in</strong> u 0 )N has tangential component −(1/ s<strong>in</strong> u 0 ) cos u 0 x u = cot u 0 (n × T).


116 SELECTED ANSWERS2.4.8 Only circles. By Exercise 2 such a curve will also have constant curvature, <strong>and</strong> byMeusnier’s formula, Proposition 2.5, the angle φ between N <strong>and</strong> n = x is constant.Differentiat<strong>in</strong>g x · N = cos φ = const yields τ(x · B) =0. Either τ =0,<strong>in</strong>which casethe curve is planar, or else x·B =0,<strong>in</strong>which case x = ±N, soτ = N ′ ·B = ±x ′ ·B =±T · B =0. (In the latter case, the curve is a great circle.)3.1.1 a. 2π s<strong>in</strong> u 03.1.3 a. 1, b.,c. 3.3.2.1 a. The semicircle centered at (2, 0) of radius √ 5; d(P, Q) =ln ( (3 + √ 5)/2 ) .3.2.4 Show first that for an arclength parametrization κ g = u ′ (s)/v(s)+θ ′ (s). Take a circleof the form α(t) = ( a + b cos t, b(s<strong>in</strong> t +1) ) <strong>and</strong> show υ(t) =‖α ′ (t)‖ =1/(s<strong>in</strong> t + 1).3.2.9 κ g = coth R3.3.4 We have κ n =II(e 1 , e 1 )=−de 3 (e 1 ) · e 1 = ω 13 (e 1 ). S<strong>in</strong>ce e 3 = s<strong>in</strong> θe 2 + cos θe 3 ,the calculations of Exercise 3 show that ω 13 = s<strong>in</strong> θω 12 + cos θω 13 , so ω 13 (e 1 ) =s<strong>in</strong> θω 12 (e 1 )=κ s<strong>in</strong> θ. Here θ is the angle between e 3 <strong>and</strong> e 3 ,sothis agrees with ourprevious result.3.3.7 We have ω 1 = bdu(<strong>and</strong> ω 2 = (a)+ b cos u) dv, so ω 12 = − s<strong>in</strong> udv <strong>and</strong> dω 12 =cos ucos u− cos udu∧ dv = −ω 1 ∧ ω 2 ,soK =b(a + b cos u)b(a + b cos u) .3.4.2 a. Tak<strong>in</strong>g ξ = f gives us ∫ 10 f(t)2 dt =0. S<strong>in</strong>ce f(t) 2 ≥ 0 for all t, iff(t 0 ) ≠0,wehavean <strong>in</strong>terval [t 0 − δ, t 0 + δ] onwhich f(t) 2 ≥ f(t 0 ) 2 /2, <strong>and</strong> so ∫ 10 f(t)2 dt ≥ f(t 0 ) 2 δ>0.3.4.9 y = 1 2 cosh(2x)A.1.1A.1.2A.2.4Consider z = x − y. Then we know that z · v i =0,i =1, 2. S<strong>in</strong>ce {v 1 , v 2 } is a basisfor R 2 , there are scalars a <strong>and</strong> b so that z = av 1 + bv 2 . Then z · z = z · (av 1 + bv 2 )=a(z · v 1 )+b(z · v 2 )=0,soz = 0, asdesired.H<strong>in</strong>t: Take u =( √ a, √ b) <strong>and</strong> v =( √ b, √ a).Let v = ∫ ba f(t) dt. If v ≠ 0, we have ‖v‖2 = v · ∫ ba f(t) dt = ∫ bav · f(t) dt ≤∫ ba ‖v‖‖f(t)‖ dt = ‖v‖ ∫ ba‖f(t)‖ dt (us<strong>in</strong>g the Cauchy-Schwarz <strong>in</strong>equality u · v ≤‖u‖‖v‖), so ‖v‖ ≤ ∫ ba‖f(t)‖ dt, asneeded.


INDEXangle excess, 79arclength, 6asymptotic curve, 47, 49, 52asymptotic direction, 47, 51, 53Bertr<strong>and</strong> mates, 20b<strong>in</strong>ormal vector, 11Bäcklund, 100C k ,1,34, 109catenary, 5catenoid, 42, 63Cauchy-Schwarz <strong>in</strong>equality, 109cha<strong>in</strong> rule, 109characteristic polynomial, 108Christoffel symbols, 54Codazzi equations, 57–60, 98compact, 58conformal, 39connection form, 99convex, 28covariant constant, 64covariant derivative, 64Crofton’s formula, 24, 32cross ratio, 93cubiccuspidal, 2nodal, 2twisted, 3curvature, 11cycloid, 3Darboux frame, 67, 98, 99developable, see ruled surface, developabledirectrix, 37Dup<strong>in</strong> <strong>in</strong>dicatrix, 53eigenvalue, 108eigenvector, 108elliptic po<strong>in</strong>t, 48Euler characteristic, 81exterior angle, 31, 79first fundamental form, 38flat, 58, 73, 80, 86, 98Frenet formulas, 11Frenet frame, 10functional, 101Gauss equation, 57, 60, 98Gauss map, 24, 43Gauss-Bonnet formula, 79, 82, 90, 99Gauss-Bonnet Theorem, 90global, 81local, 79Gaussian curvature, 48, 51, 54, 57, 78, 98constant, 59, 87generalized helix, 15geodesic, 67geodesic curvature, 68globally isometric, 71gradient, 109Gronwall <strong>in</strong>equality, 113H, 48helicoid, 35, 47, 52, 63helix, 3holonomy, 75, 78horocycle, 92hyperbolic plane, 87Kle<strong>in</strong>-Beltrami model, 95Po<strong>in</strong>caré model, 94hyperbolic po<strong>in</strong>t, 48<strong>in</strong>version, 93<strong>in</strong>volute, 19117


118 INDEXisometry, 107K, 48k-po<strong>in</strong>t contact, 54knot, 25Laplacian, 62l<strong>in</strong>e of curvature, 46l<strong>in</strong>ear fractional transformation, 89locally isometric, 39mean curvature, 48meridian, 37, 50metric, 71Meusnier’s Formula, 49m<strong>in</strong>imal surface, 48, 52, 62, 104mov<strong>in</strong>g frame, 96normal curvature, 49normal field, 32normal plane, 17oriented, 80orthogonal, 107orthonormal, 107osculat<strong>in</strong>g circle, 21osculat<strong>in</strong>g plane, 17, 21osculat<strong>in</strong>g sphere, 21parabolic po<strong>in</strong>t, 48parallel, 37, 50, 64, 72, 89parallel translate, 65parametrizationregular, 1, 34parametrized by arclength, 7parametrized curve, 1pedal property, 106planar po<strong>in</strong>t, 48Po<strong>in</strong>caré disk, 94positively oriented, 107pr<strong>in</strong>cipal curvatures, 46pr<strong>in</strong>cipal directions, 46, 51pr<strong>in</strong>cipal normal vector, 11profile curve, 37pseudosphere, 49regular, 1regular parametrization, 34rotation <strong>in</strong>dex, 26ruled surface, 37developable, 41, 58, 62, 73rul<strong>in</strong>g, 37second fundamental form, 45, 51shape operator, 44, 51smooth, 1, 34spherical coord<strong>in</strong>ates, 35stereographic projection, 36support l<strong>in</strong>e, 32surface, 34surface area, 40surface of revolution, 37symmetric, 44tangent <strong>in</strong>dicatrix, 24tangent plane, 38Theorema Egregium, 57, 98torsion, 11torus, 35total curvature, 24, 84total twist, 32tractrix, 5, 13triply orthogonal system, 53Tschebyschev net, 42twist, 32u-, v-curves, 34ultraparallels, 89umbilic, 48unit normal, 38unit tangent vector, 11variation, 101vector field, 64velocity, 1vertex, 29rectify<strong>in</strong>g plane, 17reflection, 92

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!