1 Motion and curve We shall deal with curves in the Euclidean plane and three dimensional space. Although we shall need only these two dimensions, we shall start with the general n-dimensional Euclidean space En. 1.1 1.1. Let I ⊆ R be an open interval. We shall understand points of I as valus of time t. A map f : I → R could be viewed as a motion whose trajectory is a curve in a “reasonable” case. To employ calculus we shall need in particular differentibility of the map f. We recall the real function ϕ : I → R is of the class Cr if it has continuous derivatives of the order ≤ r on I. Choosing cartesian coordinates on En, then f(t) = (f1(t), . . . , fn(t)) is n-tuple of real functions. We say f is of the class Cr if all fucntions f1, . . . , fn are fo the class Cr. One needs to show that this notion does not depend on the choice of the coordinate system. This is indeed true but a direct verification (based on a transformation from one coordinate sysytem to another one) is difficult. It is much easier to control independence only on the choice of the origin of coordinates. The choice of the origin identifies En with its associated vector space Z(En) which is an n-dimensional Euclidean vector space. 1.2 1.2. We shall therefore focus on the n-dimensional Euclidean vector space V first. We denote by ||u|| the norm of the vector u and by (u, v) the vector porduct of vectors u and v. Definition. A map v : I → V is called a vector function on the interval I. 1.3 1.3. Notion of the limit of a vector function is introduced analogoulsu as the limit of a real function. Definition. Vector function v has the limit v0 at the point t0 ∈ I if for each > 0 there exists δ > 0 such that the following holds: if t = t0 satisfies |t − t0| < δ then ||v(t) − v0|| < . We write v0 = limt→t0 v(t). If v(t0) = limt→t0 v(t) we say the vector function v is continuous at the point t0. de1.4 1.4 Definition. If the limit lim t→t0 v(t) − v(t0) t − t0 = lim t→t0 1 t − t0 v(t) − v(t0) , it is called derivative of the vector function v(t) at the piont t0. 1 We shall denote this derivative by dv(t0) dt or v (t0). Higher order derivatives are defined by the usual iteration. 1.5 1.5. Let e1, . . . , en is a basis of V . For each t ∈ I we have v(t) = v1(t)e1 + . . . + vn(t)en. Real functions vi(t), i = 1, . . . , n are called components of the vector function v(t). The following theorem has a simple proof which however belongs to calculus. Therefore we do not state it. Theorem. A vector function is continuous if all its components are continuous. The vector function v(t) has the derivative at the point t0 if all components have derivative at the point t0. Then it holds dv(t0) dt = dv1(t0) dt , . . . , dvn(t0) dt . A similar statement holds also for limits and higher order derivatives. diffscalar1.6 1.6. Now we shall state an auxiliary result which we shall need later. Let v(t) and w(t) be two vector functions of the class C1 on I. Their scalar product v(t), w(t) is a real function of the class C1 on I. Also scalar products dv dt , w and v, dw dt are real function on I. Theorem. The following holds d(v, w) dt = dv dt , w + v, dw dt . Proof. In coordinates we have v(t), w(t) = v1(t)w1(t) + . . . + vn(t)wn(t). Thus using the chain rule we have d(v(t), w(t)) dt = dv1 dt w1 + v1 dw1 dt + . . . + dvn dt wn + vn dwn dt . This is the coordinate form of our statement. 1.7 1.7. Let us consider En with its associated vector space V . Let us choose the origin P ∈ En. Then the map f : I → En determines the vector function −→ Pf : I → V , −→ Pf(t) = −−−→ Pf(t) which is called radius (or radious vector) of f. 2 Definition. The map f : I → En is called motion in the space En. We say f is motion of the class Cr, if −→ Pf is a vector function of the class Cr. Beside the word “motion” we can equivalently say path. The terminology “motion” is more illustrative, “path” has a more technical nature. The vector d −→ Pf dt does not depend on the choice of the origin. Indeed, for another point Q ∈ En we have −→ Pf = −−→ PQ + −→ Qf where −→ Pf is a constant vector. Thus d −→ Pf dt = d −→ Qf dt . de1.8 1.8 Definition. The vector d( −→ Pf) dt =: df dt is called the velocity vector of the motion f. It will be also denoted by f . At the second order we put f = (f ) ; here we already differentiate the vector function f (and analogously in higher orders). If f(t) = f1(t), . . . , fn(t) is the coordinate expression of the motion f, we have dkf(t) dtk = dkf1(t) dtk , . . . , dkfn(t) dtk . 1.9regularde1.9 1.9 Definition. The motion f : I → En of the class is called regular, if df(t) dt = o for every t ∈ I. The point of the paramter t0 at which df(t0) dt = o is called a singular point of the motion f. Here o denotes the zero vector of the space V = Z(En). We shall show two examples: (i) In the case of constant motionu f(t) = Q ∈ En, it holds for every t ∈ I that df(t) dt = o. Thus for every vaule of time t ∈ I we obtain a singular point. (ii) Consider motion x = t2, y = t3 in En, t ∈ (−∞, ∞). This moves along so colled semicubic parabola y2 − x3 = 0. We have f(t) = (t2, t3), f (t) = (2t, 3t2) hence f (0) = o. The singular point bod t = 0 is so called edge, see the picture. JS: missing picture repar1.10 1.10. Consider another open interval J with the variable τ and a bijective correspondence ϕ: J → I (i.e. a real function) of the class Cr such that dϕ dτ = 0 for every τ ∈ J. Lemma. If f : I → En is the regular motion of the class Cr then f ◦ϕ: J → En is also the regular motion of the class Cr. 3 Proof. We have d(f◦ϕ) dτ = df dt dϕ dτ where dϕ dτ is a scalar and d(f◦ϕ) dτ and df dt are vectors. Indeed, the coordinate expression of f ◦ϕ is f1(ϕ(τ)), . . . , fn(ϕ(τ)). The differentiation with respect to τ means to differentiate, at every component, a composed function with the same inner factor ϕ(τ). Thus d(f◦ϕ) dτ = df1 dt dϕ dτ , . . . , dfn dt dϕ dτ = df dt dϕ dτ . Since df dt is a nonzero vector and each dϕ dt is a nonzero scalar, also every d(f◦ϕ) dτ is a nonzero vector. de1.11 1.11 Definition. Motion f : I → En is called simple, if f is an injective mapping, i.e. for t1, t2 ∈ I, t1 = t2 we have f(t1) = f(t2). From the geometric point of view, f is a motion without self-intersections. simplecurvede1.12 1.12 Definition. The set C ⊂ En is called a simple curve of the classs Cr, if there is a simple regular motion f : I → En of the class Cr, such that C = f(I). The map f : I → En is called parametrization of the simple curve C. and the map ϕ from 1.10 is called reparametrization of the curve C. 1.13 1.13. Let J be another interval with the variable τ and g: J → En be another parametrization of the curve C of the class Cr. The rule f ϕ(τ) = g(τ) deteremines a map ϕ: J → I, t = ϕ(τ), see the picture. Theorem. ϕ je funkce tdy Cr a plat dϕ dτ = 0 pro vechna τ ∈ J. Proof. Considering coordinate epxressions f(t) = f1(t), . . . , fn(t) , g(τ) = g1(τ), . . . , gn(τ) , the functiion ϕ is determined by relations e1e1 (1) fi(t) = gi(τ) , i = 1, . . . , n . Consider an arbitrary point τ0 ∈ J and put t0 = ϕ(τ0) ∈ I. Since df(t0) dt = o, at least one of components, say k, of this vector is nonzero. Let us write the relation fk(t) = gk(τ) as e2e2 (2) fk(t) − gk(τ) = 0 . The left hand side is a function fo two variables t a τ of the class Cr on the product I ×J. We have dfk(t0) dt = 0 hence we can apply the implicit fucntion theorem to the eqaution (2) Therefore, t is determined as a function of the class Cr with the variable τ on an open neighbourhood of the point t0. At every point we find that t = ϕ(τ) is a function of the class Cr. Accoring 4 to the geometric situation, this function satisfies all equations in (1), i.e. g(τ) = f ϕ(τ) . By differentiation we find euality of vectors dg dτ = df dt dϕ dτ , where dϕ dτ is a scalar. Here dϕ(t0) dτ = 0 at some point would mean dg(τ0) dτ = o, which is a contradiction with our assumptions. We have shown that every two of parametrizations of of the class Cr a simple curve differ by a reparametrization in the sense of 1.10 and 1.12. globalcurve1.14 1.14. Now we shall introduce the notion of a global curve. Definition. A subset C ⊂ En is called curve of the class Cr, if at each point p ∈ C there is its neighbourhood U such that C ∩U is a simple curve of the class Cr. A parametrization of the intersection is called local parametrization of the curve C. 1.15 1.15. Agreement. Henceforth we shall assume the class r of the curve we consider is sufficiently high for all performed operations. This will not be usually explicitly stated. 1.16 1.16. We shall she several example in the plane E2. a) Parabola is a global simple curve. b) Circle is a curve but not a simple curve. c) This shape “quarterfoil” is a curve in our (i.e. differential geometric) definition. d) Two circles with the same center can be considered as one curve (connectivity is not assumed in the definition 1.14. It is in fact often useful to say that the border of the annulus determined by these two circles is one curve. JS: fix translation On the other hand, the whole semicubuc parabola from 1.9, the Descart curve from in e) or the lemniscata in f) are not curves in our sense. JS: missing picture de1.17 1.17 Definition. Two parametrizations f(t) and g(τ) of a simple curve C are called corresponding each other, if dϕ dτ > 0 where ϕ is the function from 1.10. Two congrunet parametrizations determine the same orientation os a simple curve C. The choice of an orientation of C thus means to determine the “direction of motion’. This is possible to do in two ways. de1.18 1.18 Definition. Let f : I → En be a local parametrization of the curve C v En. The line determined by the point f(t0), t0 ∈ I and the vector f (t0) is called tangent line of C at the point f(t0). 5 This definition is independent on the choice of parametrization, because accoridng to 1.10, two different parametrizations determine colinear vectors. Thus the tangent line at the point f(t0) has a parametrization f(t0) + v, f (t0) , v ∈ R . de1.19 1.19 Definition. The deviation of curves C and ¯C at the intersection point p is the angle of their tangent lines at this point. contactde1.20 1.20 Definition. We say two curves C and ¯C of the class Cr at the intersection point p have the contact of the order k, k ≤ r, if there are local parametrizations f(t) and ¯f(t) of, respectively, of C and ¯C on the same interval I with f(t0) = ¯f(t0) = p, if e3e3 (3) dif(t0) dti = di ¯f(t0) dti for all i = 1, . . . , k . That is, both curves “agree up to the order k” in such parametrization at the given point. contactequivpo1.21 1.21 Remark. It is easy to verify that the “contact of the order k of two curves” is an equivalnce relation. ve1.22 1.22 Theorem. Two curves C and ¯C have, at the intersection point p, the contact of the first order, if and only if their tangent lines at the point p are equal. Proof. If C and ¯C have the contact of the first order at the point p, there will exist parametrizations f(t) a ¯f(t), f(t0) = ¯f(t0) = p such that df(t0) dt = d ¯f(t0) dt . Thus their tangent lines are equal. In the opposite direction, consider ¯C with an arbitrary parametrization ¯f(¯t), f(t0) = ¯f(t0). If both tangent lines are equal then d ¯f(t0) d¯t = k df(t0) dt , k = 0. Let us perform the reparametrization ¯t = t0 + 1 k (t−t0) of ¯C. Then using the new parametrization ¯f t0 + 1 k (t − t0) of the curve ¯C, we have d ¯f(t0) dt = d ¯f(t0) d¯t · d¯t dt = d ¯f(t0) d¯t 1 k . This is equal to df(t0) dt according to the definition of k. Thus C a and ¯C have the contact of the first order. du1.23 1.23 Corollary. Tangent line is the only line which has the contact of the first order with the given curve. de1.24 1.24 Definition. The point p ∈ C is called inflection point of the curve C, if the tangent line at phas the contact of the 2nd order with the curve C. 6 inflectionve1.25 1.25 Theorem. Let f be a local parametrization of the curve C. Then p = f(t0) is the inflection point if and only if the vector d2f(t0) dt2 is colinear with the vector df(t0) dt . Proof. Put v = df(t0) dt . An arbitrary motion along the tangent line has the form g(t) = p + h(t)v where h is a real function. We have dg(t0) dt = dh(t0) dt v, d2g(t0) dt2 = d2h(t0) dt2 v which are colinear vectors. If C and its tangent line have the contact of the 2nd order, also vectors df(t0) dt a d2f(t0) dt2 are colinear. In the opposite direction, let d2f(t0) dt2 = kdf(t0) dt . Consider the parametrization of the tangent line g(t) = p + (t − t0) + k 2 (t − t0)2 v . Then dg(t0) dt = vdf(t0) dt , d2g(t0) dt2 = kv = d2f(t0) dt2 . Thus the curve C has the contact of the 2nd order with the tangent line. arclengthde1.26 1.26 Definition. A parameter s of the parametrization f : I → En of the curve C is called arc-length, if df ds = 1 for all s ∈ I. Thus the arc-length denotes “motion with constant norm of the velocity” along the curve. Let f(t) be a parametrization of the curve C. We want to find a reparametrization s = s(t) with the inverse function t = t(s) such that s is the arc. That is, 1 = df ds = df dt dt ds . Thus ds dt = df dt . Assuming parameters s and t are corresponding each other, we have ds dt = df ds = df1 dt 2 + · · · + dfn dt 2 . This means e4e4 (4) ds = (f1)2 + · · · + (fn)2 dt . Now we shall find the arc by integration. Thus the arc-length is given on every simple curve up to an additive constant. The display (4) shows that our notion of the arc-legth agrees with the length of the curve as introduced in calculus. It also agrees with the physical 7 meaning in the sense that if we move along a curve with the unique velocity then the length of the curve agrees with the length of the corresponding time interval. arcinflectionve1.27 1.27 Theorem. Assuming f(s) is the arc-length parametrization, f(s0) is the inflection point if and only if d2f(s0) ds2 = o. Proof. The vector df ds is unit which equivalentnly means df ds , df ds = 1 . Accoridng to the theorem 1.6, the differentiation of this relation yields 2 df ds , d2f ds2 = 0. That is, the vector d2f(s0) ds2 is perpendicular to the unit vector df(s0) ds . These two vector must be colinear at inflection points accoridng to the theorem 1.25. Thus d2f(s0) ds2 = o. ve1.28 1.28 Theorem. The simple curve C where all point are inflection, points is a part of a line. Proof. Considering the arc-length parametrization f(s) of the curve C, all points are inflection points if and only if d2f ds2 = o. By integration we obtain df ds = a for a constant vector a. vektor. One more integration yields f = as + b where b is another constant vector. This is a parametrization of a line. 8 2 Plane curves 2.1 2.1. Let us fix cartesian coordinates (x, y) in E2. Parametrization of curves has the form f(t) = f1(t), f2(t) , df dt = o. In particularm the graph of teh function y = f(x), x ∈ (a, b) of the class Cr is a curve of the class Cr. Its parametrization is g(t) = t, f(t) , t ∈ (a, b), thus dg dt = 1, df dt = o. We term this an explicit expression of a plane curve. 2.2 2.2. Recall that a function f : U → R of two variables defined on an open set U ⊂ R2 is of the class Cr if it has continuous partial derivatives on U of all orders ≤ r. Theorem. Let U ⊂ R2 be an open set and F : U → R be a function of the class Cr. Assume the set C defined by F(x, y) = 0 is nonempty and satisfies ∂F(x0, y0): = ∂F(x0,y0) ∂x , ∂F(x0,y0) ∂y = o for all (x0, y0) ∈ C. Then the curve C is of the class Cr. Proof. Let F(x0, y0) = 0 and assume that e.g. ∂F(x0,y0) ∂y = 0. Then according to the implicit function theorem, the set C can be locally expressed in the form y = f(x) where f(x) is a function of the class Cr. This is a local parametrization of the curve C. If ∂F(x0,y0) ∂x = 0 we can (again using the implicit function tehorem) express C locally in the form x = g(y). Definition. The point (x0, y0) such that F(x0, y0) = 0, ∂F(x0,y0) ∂x = 0, ∂F(x0,y0) ∂y = 0 is called a singular point of the set F(x, y) = 0. 2.3. Examples. (i) Consider the set x2 + y2 = a, a ∈ R, i.e. F(x, y) =pr2.3 x2 + y2 − a. The set F(x, y) = 0 is empty for a < 0 and it is a single point for a = 0 where both partial derivatives ∂F ∂x = 2x, ∂F ∂y = 2y are zero. Of course, this point is not a curve. The case a > 0 corresponds to the circle with the center at the origin and the radius √ a. The vector ∂F = (2x, 2y) is then nonzero at all points. (ii) Consider the Descrat list F(x, y) = x3 + y3 − 3axy = 0. We have ∂F = (3x2 − 3ay, 3y2 − 3ax). Assuming a = 0, (0, 0) is the unique single point. Assuming a = 0, one easily verifies that the equation ∂F = 0 has two solutions: (0, 0) and (a, a). The point (a, a) is not on the curve thus (0, 0) the unique singular point. (iii) In the case of the semicubic parabola we have F(x, y) = y2 − x3 = 0 m me ∂F = (−3x2, 2y). Thus the origin is the unique singular point. 9 implicittangentve2.4 2.4 Theorem. The tangent line to the curve F(x, y) = 0 at the point(x0, y0) is given by the equation (1) ∂F(x0, y0) ∂x (x − x0) + ∂F(x0, y0) ∂y (y − y0) = 0 . Proof. Let f1(t), f2(t) be a parametrization of this curve with f1(t0), f2(t0) = (x0, y0). Differentiatin of F f1(t), f2(t) = 0 and putting t = t0 we obtain ∂F(x0, y0) ∂x df1(t0) dt + ∂F(x0, y0) ∂y df2(t0) dt = 0 . The vector ∂F(x0, y0) is thus perpendicular to the vector df(t0) dt . The equation (2.4) describes the line through the point (x0, y0) which is perpendicular to the vector ∂F(x0, y0), i.e. the tangent line. The condition ∂F(x0, y0) = o for a curve given by the equation, thus guarantees existence of the tangent line similarly as the condition df(t0) dt = o for the parametrization of a curve. There might not be a uniue tangent line at singular points. The line through a point on the curve which is perpendicular to the tangent line at this point is called normal line. The vector ∂F(x0, y0) thus yields the direction of the normal line. 2.5implparcontact 2.5. According to 1.20, two plane curves C and ¯C have, at a intersection point p, the contact of the kth order if there exist local parametrizations f1(t), f2(t) and ¯f1(t), ¯f2(t) of, respectively, C and ¯C on the same interval I such that parcontactparcontact (2) dif1(t0) dti = di ¯f1(t0) dti , dif2(t0) dti = di ¯f2(t0) dti , i = 1, . . . , k, where t0 is the parametr of the intersection point p. The direct approach to the question whether such parametrizations do or do not exist is rather complicated in general. However, there is a very simple answer if ¯C is given by the equation F(x, y) = 0. In this case we can form the one variable function (3) Φ(t) = F f1(t), f2(t) . Theorem. Curves C ≡ f1(t), f2(t) and ¯C ≡ F(x, y) = 0 have, at a intersection point (x0, y0) = f1(t0), f2(t0) the contact of the kth order if and only if implcontactimplcontact (4) diΦ(t0) dti = 0 , i = 1, . . . , k . 10 Proof. Let ¯f1(t), ¯f2(t) be a local parametrization of the curve ¯C such that the condition (2) for the contact is satisfied. Then (5) F ¯f1(t), ¯f2(t) = 0 for all t, i.e. all derivatives of the composed function of the left hand side are zero. Also the function Φ is composed with outer factor F(x, y) and inner factors f1(t) and f2(t). According to our assumption about the contact, the derivatives up to the order k of inner factors at the point t0 are the same for both ¯f1(t) and ¯f2(t). Thus (4) holds. In the opposite direct, assume (4) holds. Further assume ∂F(x0,y0) ∂y = 0. We shall locally parametrize the curve ¯C in the form f1(t), g(t) where the function g(t) is determined by (6) F f1(t), g(t) = 0 . This is always possible. Indeed, consider G(t, y) = F f1(t), y , which is well defined on some neighbourhood V of the point (t0, y0). We have ∂G(t0, v0) ∂y = ∂F(x0, y0) ∂y = 0 , thus we can use the implicit function theorem for the function G(t, y) = 0. We need to show that f2gcontactf2gcontact (7) dif2(t0) dti = dig(t0) dti , i = 1, . . . , k . On V × R we shall consider the function H of three variabls, H(t, y, z) = G(t, y) − z . It satisfies H(t0, y0, 0) = 0 and ∂H(t0,y0,0) ∂y = ∂G(t0,y0) ∂y = 0. Hence using the implicit function theorem once more, from H = 0 we can locally (and uniquelly) compute y = K(t, z). Since G t, g(t) = 0 and G t, f2(t) = Φ(t), we have g(t) = K(t, 0) a f2(t) = K t, Φ(t) . Similarly as in the first part of the proof, we have th same outer factor K(t, z). Derivations of the constant function t → 0 and the function Φ(t) up to the order k at the point t0 agree (because they are zero). According to the chain rule, (7) implies (4) 11 2.6 2.6. Now we shall discuss how to approximate an arbitrary plane curve C at a given point p using circles. Definition. A circle which has the 2nd order contact with the curve C at the point p ∈ C, is called osculating circle at the point p. ve2.7 2.7 Theorem. Assume p ∈ C is not an inflection point. Then there is exactly one osculating curve at p. Proof. Denote by (a, b) the center and by r the radius of the circle, i.e. its equation is ocircleocircle (8) (x − a)2 + (y − b)2 − r2 = 0 . Using the theorem 2.5 we shall find a condition for (8) to have the 2nd order contact with the curve given by the parametrization f1(t), f2(t) at the point t0. We have Φ(t) = f1(t) − a 2 + f2(t) − b 2 − r2 , Φ (t) = 2(f1 − a)f1 + 2(f2 − b)f2 , Φ (t) = 2(f1)2 + 2(f1 − a)f1 + 2(f2)2 + 2(f2 − b)f2 . Coordinates a, b are solutions of the equation Φ = 0, Φ = 0 which are equivalent to osystemosystem (9) af1 + bf2 = f1f1 + f2f2 , af1 + bf2 = f1f1 + f2f2 + f 2 1 + f 2 2 . Away from inflection points, vectors (f1, f2) and (f1 , f2 ) are linearly independent. Hence the determinant of the system (9) is nonzero and these equations determine the unique pair (a, b). The radius r is then given by the equation Φ = 0. ve2.8 2.8 Theorem. The radius r of the osculationg curve satisfies oradiusoradius (10) r2 = (f 2 1 + f 2 2 )3 (f1f2 − f2f1 )2 Proof. One computes (using e.g. the Cramer’s rule) from (9) that a = f1 − f2(f 2 1 + f 2 2 ) f1 f2 f1 f2 , b = f2 + f1(f 2 1 + f 2 2 ) f1 f2 f1 f2 . The relation r2 = (f1 − a)2 + (f2 − b)2 then yields (10). 12 2.9 2.9. Osculating curves do not exist at inflection points. The tangent line has the 2nd order contact with the curve hence this curve would have to have the contact of the 2nd order with the osculating curve accoring to 1.21. However, a simple computation reveals that a circle has the contact of the 1st order with its tangent line. Indeed, we can choose such coordinate system such that the circle is given by the parametrization x = r cos t, y = r sin t. Its tangent line at the point t = 0 has the equation x − r = 0. We have Φ(t) = r cos t − r thus Φ(0) = 0, Φ (0) = −r sin 0 = 0 but Φ (0) = r cos 0 = 0. JS: missing picture planecurvaturede2.10 2.10 Definition. Let r be the radius of the osculating curve at the point p ∈ C (away from inflection points). The number κ = 1 r is called curvature of the curve C vat the point p. We define κ = 0 at inflection points. This terminology is motivated by the observation that a smaller radius of the osculating curve means the curve is more “curved”. The center of the osculating curve is called center of the curvature of the curve C at the given point. de2.11 2.11 Definition. Assume p ∈ C is not the inflection point. If the osculating curve at p ∈ C has the contact of the 3rd order with C the p is called vertex of the curve. We shall show in 2.16 that in the case of an ellipse, our general notion of vertices agrees with the usual definition of vertices of ellipses. The oscullating curve at vertex of a curve is called hyperosculating. arccurvature2.12 2.12. Consider the arc-length parameter s. Then the vector e1 = df ds is unit and perpendicular to the vector de1 ds = d2f ds2 according to 1.26 and the proof the theorem 1.27. Here the inflection point f(s0) is characterized by de1(s0) ds = o. At the point f(s0) (away from inflection points) we denote by e2(s0) the unit vector parallel with de1(s0) ds in the same direction. Thus e1(s0) and e2(s0) is a pair of orthonormal vectors. Theorem. We have de1(s0) ds = κ(s0) and the center of the osculating circle lies on the halfline given by the point f(s0) and the vector e2(s0). Proof. One can derive this from the expression for the center of the osculating curve in (10). But it will be useful for later considerations to perform 13 the whole computation once more (with some simplifications). Since the osculating circle has the contact of the 1st order with the tangent line, its center lies on the normal line. Thus this center is of the form f(s0)+re2(s0) for some r ∈ R. The equation of the circle with this center and the radius r can be written using the scalar product as z − f(s0) − re2(s0), z − f(s0) − re2(s0) − r2 = 0 , where z = (x, y) is an arbitrary point of the plane. For the computation of the contact we shall therefore use the function Φ(s) = f(s) − f(s0) − re2(s0), f(s) − f(s0) − re2(s0) − r2 . By the differentiation and using 1.6 we obtain 1 2 dΦ ds = e1(s), f(s) − f(s0) − re2(s0) . Conditions Φ(s0) = 0 and dΦ(s0) ds = 0 are satisfied; geometrically this follows from the fact that we chose the center on the normal line. One more differentiaion yields e2.11e2.11 (11) 1 2 d2Φ ds2 = de1(s) ds , f(s) − f(s0) − re2(s0) + e1(s), e1(s) . This must be zero at the point s0 hence e2.12e2.12 (12) r de1(s0) ds , e2(s0) = 1 . Since the vector de1(s0) ds is congruently parallel with the vector e2(s0), the scalar product (12) is equal to the norm of this vector. Our statement is thus a direct conseuqence of the definition 2.10. e12curvaturedu2.13 2.13 Corollary. We have de1(s) ds = κ(s)e2(s). Proof. It follows from the theorem 2.12 away from inflection points and from the theorem 1.27 in inflection points. ve2.14 2.14 Theorem. We have de2(s) ds = −κ(s)e1(s). Proof. Since e2 is a unit vector, we have (e2, e2) = 1. By differentiation we obtain e2, de2 ds = 0. Thus the vector de2 ds is perpendicular to e2, i.e. 14 de2 ds = ce1. Since vectors e1 and e2 are perpendicular, we have (e1, e2) = 0. By differentiation we obtain 0 = de1 ds , e2 + e1, de2 ds = κ + c . ve2.15 2.15 Theorem. The point f(s0) is the vertex if and only if dκ(s0) ds = 0. Proof. We shall continue in the proof of the theorem 2.12 and use also the corollary 2.13. We obtain 1 2 d2φ ds2 = κ(s) e2(s), f(s) − f(s0) − re2(s0) + 1 . Further differentiation yields the condition of the contact of the 3rd order 0 = 1 2 d3φ(s0) ds3 = dκ(s0) ds · (−r) + κ(s0) − κ(s0)e1(s0), −re2(s0) . Our statement now follows from e1(s0) ⊥ e2(s0) and r = 0. vrchol2.16 2.16 Corollary. Consider an arbitrary parametrization f(t) of the curve C. Then away from inflection points, the point f(t0) is vertex if and only if dκ(t0) dt = 0. Proof. The transformation from t to s is realized using a reparametrization t = ϕ(s), t0 = ϕ(s0). It follows form the chain rule that dκ(ϕ(s0)) ds = dκ(t0) dt dϕ(s0) ds . Here dϕ(s0) ds = 0 since ϕ is a reparametrization. From this in particularly follows that vertex of an ellipse in the differential geometric sense are vertices of an ellipse in the classical sense since the curavture at these points achieves maximum of minimum. onlyverticesve 2.17 2.17 Theorem. A simple curve whose every point is vertex, is a part of the circle. Proof. The center of the osculating circle is c(s) = f(s) + 1 κ e2(s). If every point is vertex, κ will be a constant. Thus by differentiation and using 2.13, we obtain dc(s) ds = e1(s) − 1 κ κe1(s) = o . Thus c(s) is a fixed point and also the radius 1 κ is constant. All osculating curves thus coincide and the curve lies on a circle. 15 2.18 Definition. Relations e2.13e2.13 (13) df ds = e1 , de1 ds = κe2 , de2 ds = −κe1 are called Frenet formulae of the plane curve C without inflection points. “Moving” frame f(s), e1(s), e2(s) is called Frenet frame of the curve C. 2.19 2.19. Now we shall show how to use relations (13) in order to characterize congruence of plane curves. Definition. Curves C and ¯C ⊂ E2 are called congruent if there is an Euclidean transformation ϕ: E2 → E2 such that ϕ(C) = ¯C. ve2.20 2.20 Theorem. Let C, ¯C be curves without inflection points, f : I → E2, ¯f : I → E2, respectively their arc-length parametrizations on the same interval I and κ(s), ¯κ(s), respectively be their curvatures. Then curves C and ¯C are congruent if and only if κ = ¯κ on I. Proof. One direction is obvious: an Euclidean transformation maps arclength to arc-length and preserves the contact thus radii of osculating circles at corresponding points must be the same. In the opposite direction, consider C resp. ¯C with Frenet frame f(s), e1(s), e2(s) resp. ¯f(s), ¯e1(s), ¯e2(s) . Thus beside (13) we have also e2.14e2.14 (14) d ¯f ds = ¯e1 , d¯e1 ds = κ¯e2 , d¯e2 ds = −κ¯e1 with teh same κ. Thus (13) and (14) is the same system of differential equations for the 6-tuple of real functions which are components of f, e1 and e2. Given s0 ∈ I, the triple of vectors f(s0), e1(s0), e2(s0) as well as the triple ¯f(s0), ¯e1(s0), ¯e2(s0) is formed by the point and the pair of orthonormal vectors. Hence there exists a unique Euclidean transformation ϕ: E2 → E2 which transforms f(s0) to ¯f(s0), e1(s0) to ¯e1(s0) and e2(s0) to ¯e2(s0). Thus the parametrization ¯f : I → E2 of the curve ¯C together with vector functions ¯e1(s), ¯e2(s) and the parametrization ϕ ◦ f : I → E2 of the curve ϕ(C) together with vector functions ϕ ◦ e1, ϕ ◦ e2 satisfy the same system of differential equations with the same initial conditions. According to the theorem of the unique existence of a solution of a the system of differential equations, we have ¯f = ϕ ◦ f, ¯e1 = ϕ ◦ e1, ¯e2 = ϕ ◦ e2. The first relation ¯f = ϕ ◦ f implies ¯C = ϕ(C). pr2.21 2.21 Example. The assumption that curves C and ¯C are without inflection points, is essential. Consider curves given explicitly as y = x3 and 16 y = |x|3, x ∈ (−∞, ∞). Both curves are of the class C2 and have the same JS: missing picture curvature as a function of the arc-length. But they are not congruent. 2.22 2.22. The next statement can be briefly rephrased as that we can prescribe the curvature arbitrarily. Theorem. Let κ : I → R be a positive function. Then locally there exists a curve C parametrized by arc-length on I such that κ is its curvature. Idea of the proof: We solve the system of equations (13). Remark. Globally, this curve might not be simple. For example, if κ = 1 r is a constant, the solution of teh corresponding system of differential equations is the circle x = r cos s r , y = r sin s r . Then for s ∈ (−∞, ∞) one goes along the circle repeatedly. 2.23 2.23. We shall finish this section with a global result about plane curves. Recall the subset in E2 is called bounded if it lies inside a circle. Definition. The plane curve C of the class Cr is called oval of the class Cr if it is the border of a bounded convex set in E2. Examples: (i) (ii) JS: missing picture ve2.24 2.24. Four vertex theorem. Each oval of the class C3 without points of inflection, has at least four vertices. Proof. Consider C parametrizad by the arc-length f(s) = f1(s), f2(s) on the interval s ∈ [0, a] such that for s = a the oval closes, i.e. f(0) = f(a). Thus the curvature κ is in fact defined on the closed interval hence it reaches its maximum and minimum. This yields two vertices of the oval C. We can assume that κ ha minimum at s = 0 and it has maximum at some point b ∈ (0, a). Choose f(0) to be the origin, f(b) on the x-axis and the orientation of the y-axis such that f2(s) > 0 for s ∈ (0, b). (If this holds for one point, it holds for all points by the convexity). Then f2(s) < 0 for s ∈ (b, a) again using the convexity. The case κ(0) = κ(b) i.e. κ equal to a constant, is the circle (according to Theorem 2.17) and we can exclude this case. JS: missing picture Assume now that f(0) and f(b) are unique vertices. Then dκ ds > 0 on (0, b) and dκ ds < 0 on (b, a). The integration by parts now yields 0 < a 0 dκ ds f2ds = κf2 a 0 − a 0 κ df2 ds ds . 17 But κf2 a 0 = 0 because f(0) = f(a) and κ(0) = κ(a). Let us expand the relation vztah de1 ds = κe2. We have e1 = df1 ds , df2 ds . Since e2 is the unit vector perpendicular to e1, we have e2 = ± df2 ds , df1 ds hence d2f1 ds2 = ±κdf2 ds . Thus 0 < − a 0 κ df2 ds ds = ± a 0 d2f1 ds2 ds = ± df1 ds a 0 = 0 , because df1(0) ds = df1(a) ds according to the periodicity of the parametrization of the oval. This is a contradiction. In fact we have shown there exists another point where dκ ds changes the sign, i.e. κ has either minimum or maximum at this point. But minima and maxima appear in pairs. Thus the existence of the fourth vertex follows. 3 Envelope of a family of plane curves 3.1 3.1. Consider the one-parameter family of plane curves given by the equa- tions e3.1e3.1 (1) F(x, y, t) = 0 , t ∈ I where F(x, y, t) is a function of the class C1 defined on an open set U ⊂ R3. Let us denote by Ct0 , t0 ∈ I the curve of the equation F(x, y, t0) = 0, i.e. we consider (1) as the system of plane curves (Ct). 3.2 3.2. Intersection points of curves Ct and Cs, t = s are determined by the pair of equations F(x, y, t) = 0 , F(x, y, s) = 0 . This system of equations is oviously equivalent with the system F(x, y, t) = 0 , F(x, y, s) − F(x, y, t) s − t = 0 . Considering a fixed t, we obtain the following equation in the limit s → t, e3.2e3.2 (2) F(x, y, t) = 0 , ∂F(x, y, t) ∂t = 0 . Definition. Points determined by the equation (2) are called characteristic points on the curve Ct. The set of such points for all t ∈ I is called charakteristic set of the system (Ct). 18 From the computational point of view, we have two basic possibilities how o express the characteristic set. If we eliminate the parametr t from (2), we express the characteristic set in the form of an equation G(x, y) = 0. If we compute x and y from (2) as function of t, we obtain a parametrization of the charatcteristic set. 3.3 3.3. We say two curves touch each other in the intersection point if they have the contact of the 1st order, i.e. the same tangent line. Definition. The curve D with a parametrization f(t), t ∈ (a, b) ⊂ I is called envelope of the family (Ct) if D touches the curve Ct0 at the point f(t0) for all t0 ∈ (a, b). ve3.4 3.4 Theorem. Each envelope of the family (Ct) is a subset of its characteristic set. Proof. The condition that each point of the envelope f(t) = f1(t), f2(t) lies on the curve Ct, is e3.3e3.3 (3) F f1(t), f2(t), t = 0 . The condition that tangent lines for D and Ct coincide at the point f(t), has the form e3.4e3.4 (4) ∂F(f1(t), f2(t), t) ∂x df1(t) dt + ∂F(f1(t), f2(t), t) ∂y df2(t) dt = 0 . By differentiation of (3) we obtain e3.5e3.5 (5) ∂F(f1(t), f2(t), t) ∂x df1(t) dt + ∂F(f1(t), f2(t), t) ∂y df2(t) dt + ∂F(f1(t), f2(t), t) ∂t = 0 . Subtracting (4) from (5) yields e3.6e3.6 (6) ∂F(f1(t), f2(t), t) ∂t = 0 . Thus every envelope is a part of the characteristic set. 3.5 3.5. Consider a very simple case of the family of circles centered on the xaxis and the constant radius r. That is, F(x, y, t) = (x − t)2 + y2 − r2 = 0. Then ∂F ∂t = −2(x − t) = 0. Putting x = t to the first equation, we get y = ±r. Of course, both these lines are envelopes. 3.6 3.6. In the opposite direction, we have the following: 19 Theorem. If the curve f(t) is a solution of (2), it is an envelope of the family (Ct). Proof. The curve f(t) satisfies (3), hence f(t) ∈ Ct. By differentiation we obtain (5). Further we have (6) and subtracting (6) from (5), we obtain (4). Thus f(t) touches the curve Ct. 3.7 3.7. Having a pair of functions x = f1(t), y = f2(t) which is a solution of (2), then it is an envelope of the family (Ct) assuming further that f(t) = f1(t), f2(t) is a curve. In particular df dt = o. The picture shows first the family of curves centered on a cirle of the radius r with constant radius < r, where the inner and outer envelopes are cicrles. The second case is = r, where the inner envelopes “degenerates” to a point. JS: missing picture 3.8 3.8. Normal lines of an arbitrary plane curves C form a one-parametr family of curves. Definition. The characteristic set of the family of normal lines of the curve C is called evolute of the curve C. Theorem. Evolute of the curve C without inflection points coincides with the set of centers of their osculating curves. Proof. Let z = (x, y) be an arbitrary point in the plane. We shall parametrize C by the arc-length and consider its Frener frame e1(s), e2(s) at the point f(s). The equation of the normal line at the point f(s) then is e3.7e3.7 (7) F(x, y, s) = e1(s), z − f(s) = 0 . Using Frenet formulae. we find the condition e3.8e3.8 (8) ∂F ∂s = κ(s)e2(s), z − f(s) − e1(s), e1(s) = 0 . The characteristic set is solution of equations (7) and (8), which we shall find geometrically. It follows from (7) that plyne z = f(s) + c(s)e2(s) . Putting this into (8), we get κ(s)c(s) − 1 = 0, thus c(s) = 1 κ(s) . This is the center of the osculating circle. 20 3.9. Above we found the parametric expression of the evolute, z(s) = f(s) + 1 κ(s) e2(s) . Thus dz ds = e1(s) − κ (s) κs2 e2(s) − e1(s). If κ (s) = 0, this vector is nonzero. This means, that in some neighbourhood of the point, which is not a vertex, is evolute a cirve. The picture shows the evolute of an ellipse. Their edges correspond to JS: missing picture vertices of the ellipse. 4 Spacial curves and surfaces Beside a parametric expression, a spacial curve can be also given as an intersection of two surfaces. In the study of spacial curves, we shall use their contact with certain auxiuliary surfaces. Now we shall give a general definition of surfaces in E3. 4.1 4.1. We shall need notion of vector functions of two variables. To simplify the rpesentation, we shall work only with 3-dimensional Euclidean vector space V . Coordinates of the point u ∈ R2 will be denoted by (u1, u2). Let D ⊂ R2 be an open set. The mapping w: D → V will be called vector function of two variables. If e1, e2, e3 is a basis of V , we have w(u) = w(u1, u2) = w1(u1, u2)e1 + w2(u1, u2)e2 + w3(u1, u2)e3. Real function w1, w2, w3 are called components of the vector function w and we write e4.1e4.1 (1) w(u1, u2) = w1(u1, u2), w2(u1, u2), w3(u1, u2) . The limit and continuity of vector functions of the vector function w are defined similarly as in 1.3. We say w has the limit v ∈ V at the point u0 = (u0 1, u0 2), if for each ε > 0 there is δ > 0 such that |u1 −u0 1| < δ, |u2 −u0 2| < δ, (u1, u2) = (u0 1, u0 2) implies w(u1, u2) − v < ε. We write lim u→u0 w(u) = v. Further, w is continuous at the point u0 if lim u→u0 w(u) = w(u0). 4.2 4.2. Partial derivatives of the vector fucntion w are defined by ∂w(u0) ∂u1 = lim u1→u0 1 w(u1, u0 2) − w(u0 1, u0 2) u1 − u0 1 , ∂w(u0) ∂u2 = lim u2→u0 2 w(u0 1, u2) − w(u0 1, u0 2) u2 − u0 2 Higher order partial derivatives ∂kw ∂ui 1∂uj 2 , i + j = k, are defined by the usual iteration. 21 As in 1.5, a vector function is continuous if and only if all its components are continuos. An anlogousl statement holds also for limits and partial derivatives. In particular, following (1) we have ∂1w: = ∂w ∂u1 = ∂w1 ∂u1 , ∂w2 ∂u1 , ∂w3 ∂u1 , ∂2w: = ∂w ∂u2 = ∂w1 ∂u2 , ∂w2 ∂u2 , ∂w3 ∂u2 and similarly for higher order partial derivatives. We say the function w: D → V is of the class Cr, if it has continuous partial derivatives of the order ≤ r at the point D. 4.3 4.3. Consider V as the associate vector space of E3. Choose an auxiliary origin P ∈ E3. Then the mapping f : D → E3 determines radius vector which is the vector function −→ Pf : D → V , −→ Pf(u) = −−−−→ Pf(u). We put e4.2e4.2 (2) ∂1f = ∂f ∂u1 = ∂( −→ Pf) ∂u1 , ∂2f = ∂f ∂u2 = ∂( −→ Pf) ∂u2 . Similarly as in 1.7, this does not depend on the choice of the origin P. Here (2) are vector functions of two variables. By iteration we have e4.3e4.3 (3) ∂11f = ∂2f ∂u1∂u1 , ∂12f = ∂2f ∂u1∂u2 , ∂22f = ∂2f ∂u2∂u2 and similarly for higher orders. de4.4 4.4 Definition. The set S ⊂ E3 is called simple surface of the class Cr, if there is ana open set D ⊂ R2 and an injective mapping f : D → E3 of the class Cr such that S = f(D) and vectors ∂1f a ∂2f are linearly independent at each point of the set D. JS: missing picture We say f is parametrization of the surface S and D is parameter space. The condition that vectors ∂1f and ∂2f are linearly indepedent shall be written in the form ∂1f × ∂2f = o where × denotes the vector product. We shall illustrate the meaning of this condition on a parametrization of the plane E3. Consider D = R2 and put f = P + u1a + u2b , P ∈ E3 , a, b ∈ V, u1, u2 ∈ R . In coordinates (x, y, z) on E3 we have x = p1 + u1a1 + u2b1 , y = p2 + u1a2 + u2b2 , z = p3 + u1a3 + u2b3 . 22 Then ∂1f = a a ∂2f = b. From the analytic geometry we know that f determines the plane if and only if vectors a and b are linearly independent. If these vector are linearly dependent, we obtain only a line, and the case a = b = o yiedls only a point. 4.5 4.5. Given the fucntion z = f(x, y) tdy Cr of two variables on D ⊂ R2 then its graph ¯f(x, y) = x, y, f(x) , ¯f : D → R3 is a simple surface of the class Cr. The reason is that ∂1 ¯f = 1, 0, ∂f ∂x , ∂2 ¯f = 0, 1, ∂f ∂y and these vectors are linearly independent everywhere. We say that f(x, y) is an explicit description of the surface. de4.6 4.6 Definition. The subset S ⊂ E3 is called surface of the class Cr if for each p ∈ S there is its neighbourhood U such that U ∩ S is a simple surface of the class Cr. Examples. JS: missing picture a) Rotational paraboloid is globally a simple surface. b) The sphere JS: check the translation is a surface which is not simple. c) Anuloid is the surface given by the rotation of the circle around the axis which lies in the same plane and has empty intersection with the cicrle. The physical model is “pneumatika”. JS: check the translation d) Also “an v lec” is an interesting global example of a surface. 4.7 4.7. Agreement. Further we shall assume that the class r of the the suface or function under consideration is high enough for required constructions and this will not be usually explicitly stated. 4.8 4.8. A curve on a surface will be usually given in the parameter space D, i.e. u = u(t), tj. u1 = u1(t), u2 = u2(t), t ∈ I. On the surface S = f(D) then we have the curve f u(t) = f u1(t), u2(t) . Theorem. Tangent lines of all curves on the surface S at the point pinS fill the plane which is called tangent plane of the surface S at the point p. Proof. Let p = f(u0). The velocity vector of the motion f u(t) , u(t0) = u0 is given by the differentiation of the composed function e4.4e4.4 (4) df(u1(t0), u2(t0)) dt = ∂f(u0) ∂u1 du1(t0) dt + ∂f(u0) ∂u2 du2(t0) dt . Hence this is a linear combination of vectors ∂1f(u0) and ∂2f(u0). In the opposite direction, for arbitrary vector a = a1∂1f(u0) + a2∂2f(u0) we have the motion u(t) = u1(t), u2(t) such that du1(t0) dt = a1, du2(t0) dt = a2. Considered tangent lines this fill the whole plane given by the point p and vectors ∂1f(u0) and ∂2f(u0). 23 The tangent plane of the surafce S at the point p will be denoted by τpS and its associated vector space TpS is called tangent vector space of S at the point p. The previous theorem shows the geometric meaning of the condition of linear independence of vectors ∂1f a ∂2f which guarantees existence of the tangent plane. 4.9 4.9. Henceforth we fix the coordinate system (x, y, z), i.e. E3 ≈ R3. Theorem. Let U ⊂ R3 be an open set and F : U → R is a function of the class Cr such that the set S given by the equation F(x, y, z) = 0 is nonempty and ∂F(x0, y0, z0): = ∂F(x0, y0, z0) ∂x , ∂F(x0, y0, z0) ∂y , ∂F(x0, y0, z0) ∂z = o for each (x0, y0, z0) ∈ S. Then S is a surface of the class Cr. Proof. Let F(x0, y0, z0) = 0 and e.g. ∂F(x0,y0,z0) ∂z = 0. According to the implicit function theorem the equation F(x, y, z) = 0 locally yiedls z = f(x, y) for a function f of the class Cr. Locally this is an explicit description of the surfaces S. If ∂F(x0,y0,z0) ∂y = 0 resp. ∂F(x0,y0,z0) ∂x = 0, one can locally compute y = g(x, z) resp. x = h(y, z). The point (x0, y0, z0) at which ∂F(x0, y0, z0) = o is called singular point of the set F(x, y, z) = 0. 4.10. Examples. (i) The case F(x, y, z) = x2 + y2 + z2 − a is similar as4.10 in ??. The set F(x, y, z) = 0 is empty for a < 0. If a = 0, the equation is satisfied only by the origin which is the unique singular point. If a > 0, we have the sphere centered at the origin wit the radius √ a. The vector ∂F = (2x, 2y, 2z) is nonzero at all points. (ii) Consider “rotacni kuzel” F(x, y, z) = z2 − x2 − y2 = 0. The point JS: missing picture (0, 0, 0) is the unique singular point. Observe the tangent plane does not exist at this point. ve4.11 4.11 Theorem. The equation of the tangent plane of the surface S given by the equation F(x, y, z) = 0 at (x0, y0, z0) ∈ S is e4.5e4.5 (5) ∂F(x0, y0, z0) ∂x (x−x0)+ ∂F(x0, y0, z0) ∂y (y −y0)+ ∂F(x0, y0, z0) ∂z (z −z0) = 0 . 24 Proof. Let the curve f1(t), f2(t), f3(t) lie on S and goes through the point bodem (x0, y0, z0) ∈ S for t = t0. Then F f1(t), f2(t), f3(t) = 0 . Differentiating of the composed function and putting t = t0, we obtain e4.6e4.6 (6) ∂F(x0, y0, z0) ∂x df1(t0) dt + ∂F(x0, y0, z0) ∂y df2(t0) dt + ∂F(x0, y0, z0) ∂z df3(t0) dt = 0 . Thus the normal vector of the plane (5) is perpendicular to the tangent vector of any curve on S hence (5) is the tangent plane. de4.12 4.12 Definition. Consider the surface S. The line NpS through the point p ∈ S and perpendicular to the tangent plane τpS, is called normal line of the surface S at the point p. Thus the vector ∂F(x0, y0, z0) is the directional vector of the normal line of the surface F(x, y, z) = 0 at its point (x0, y0, z0). The condition ∂F = o geometrically guarantees existence of the tangent plane as well as the condition ∂1f × ∂2f = o in the case of a parametric description of the surface. 4.13 4.13. We shall study the question when the intersection of two surfaces e4.7e4.7 (7) F(x, y, z) = 0 , G(x, y, z) = 0 is a curve Theorem. Let U ⊂ R3 be an open set and F, G: U → R be functions of the class Cr such that the set C given by the equation (7) is nonempty and vectors ∂F(x0, y0, z0) and ∂G(x0, y0, z0) are linearly independent for each (x0, y0, z0) ∈ C. Then C is a curve of the class Cr. Proof. Since vector ∂F and ∂G are linearly independent, there is at least one nonzero subdeterminant of the order 2 in the matrix e4.8e4.8 (8) ∂F ∂x ∂F ∂y ∂F ∂z ∂G ∂x ∂G ∂y ∂G ∂z . If this is the subdetereminant ∂F ∂y ∂F ∂z ∂G ∂y ∂G ∂z , 25 it follows from the generalized implicit function theorem that from (7) one can locally compute y = f(x) a z = g(x). Here f and g are again function of the class Cr. (This theorem can be found at 2.6 of the textbook “´Uvod do glob´aln´ı anal´yzy” which is stated as [5] in the list of references.) Thus t, f(t), g(t) is locally a parametrization of the curve given by equations F = 0 and G = 0. If another subdeterminant of the order 2 is nonzero, we can locally express x and z as fucntions of y or x and y as functions of z. 4.14 4.14. We shall illustrate the generalized implicit function theorem on tje simplest example of two linear equations F(x, y, z) = a1x + a2y + a3z = 0 , G(x, y, z) = b1x + b2y + b3z = 0 . In this case we have ∂F ∂y ∂F ∂z ∂G ∂y ∂G ∂z = a2 a3 b2 b3 and if this determinant is nonzer, one can use the Cramer’s rule to compute y and z. 4.15 4.15. Geometrically, the theorem 4.13 says that the intersection of two surfaces S1 and S2 is locally a curve in a neighbourhood of a point p ∈ S1 ∩ S2 where the tangent planes τpS1 and τpS2 are different. A simple examples of two touching spheres (which intersect in a single point) shows that this condition is necessary. An interesting example is so called Viviani curve which is the intersection of the sphere and “v´alce” with half radius which goes through the JS: add translation JS: missing picture center of the sphere, see the view “from above” in a). The tangent planes of both surfaces are different surfaces with the exception of the point A. Indeed, here the intersection of both surfaces is not locally a curve in our sense, see the “front” point of view in b) and the general picture in c). 4.16 4.16. The definition of the contact of a curve with a plane reduces to contact of two curves. Definition. We say the curve C and the surface S have the contact of the k-th order at the intersection point p, if there exists a curve ¯C on S such that C and ¯C have the contact of the kth order at the point p. 26 One can easily see that C and S have the contact of the 1st order if and only if the tangent line of the curve lies in the tangent plane of the surface. 4.17 4.17. The following simple condition to determine the contact of a curve with a surafce is similar to the theorem 2.5. Let S be given by the equation F(x, y, z) = 0 and C is given by the parametrization f1(t), f2(t), f3(t) . Theorem. Let f(t0) = (x0, y0, z0) be an intersection point of the curve C and the surface S. Consider the function Φ(t) = F f1(t), f2(t), f3(t) . Then C and S have the contact of the order k if and only if e4.9e4.9 (9) diΦ(t0) dti = 0 , i = 1, . . . , k . Proof. Let ¯f(t) be a parametrization of the curve ¯C on the surface S such that derivatives of f(t) and ¯f(t) coincide up to the order k at t = t0. Since ¯C lies on S, we have e4.10e4.10 (10) F ¯f1(t), ¯f2(t), ¯f3(t) = 0 , hence all derivatives with respect to t of the left hand side are zero. The function Φ(t) and (10) have the same outer factor F(x, y, z) and derivatives of inner factors up to the order k coincide according to the condition of contact. Thus (9) holds. In the opposite direction, let e.g. ∂F(x0,y0,z0) ∂z = 0. According to the implicit function theorem, the equation F f1(t), f2(t), z = 0 locally determines the function z = g(t) and the curve ¯C ≡ f1(t), f2(t), g(t) lies on S. It is sufficient to show e4.11e4.11 (11) dif3(t0) dti = dig(t0) dti , i = 1, . . . , k . Put G(t, z) = F f1(t), f2(t), z . This function is defined on some neighborhood V of the point (t0, z0). Consider the function of three variables e4.12e4.12 (12) H(t, z, w) = G(t, z) − w . on V ×R. According to the implicit function theorem, the equation H(t, z, w) = 0 locally allows to compute z = K(t, w). Since G t, g(t) = 0 and G t, f3(t) = Φ(t), we have g(t) = K(t, 0) a f3(t) = K t, Φ(t) . As in the proof of the theorem 2.5, we obtain (11). 27 5 Frenet frame of spacial curves Consider a curve C ⊂ E3. ve5.1 5.1 Theorem. There exists a unique plane ω at non-inflection p ∈ C which has the contact of the 2nd order with C. Then ω is called osculating plane of the curve C at the point p. Proof. Consider a parametrization f(t) = f1(t), f2(t), f3(t) of the curve C and an arbitrary plane ax + by + cz + d = 0. According to 4.17 we consider the function Φ(t) = af1(t) + bf2(t) + cf3(t) + d . For the 2nd order contact we have conditions Φ(t0) = 0 and dΦ(t0) dt = a df1(t0) dt + b df2(t0) dt + c df3(t0) dt = 0 , d2Φ(t0) dt2 = a d2f1(t0) dt2 + b d2f2(t0) dt2 + c d2f3(t0) dt2 = 0 . These conditions mean that the normal vector (a, b, c) of the required plane is tangent to vectors df(t0) dt and d2f(t0) dt2 . Since these two vectors are linearly independent, the reuired plane is unique. Similarly as the osculating circle of a plane curve, the condition of the 2nd order contact of the osculating plane with the spacial curve means that the osculating plane approximates the curve in the best way (among all possible planes). du5.2 5.2 Corollary. At a non-inflection point f(t0), the associated vector space of the osculating plane is given by vectors df(t0) dt a d2f(t0) dt2 . Thus its equation can be epxressed in the form x − f1(t0), y − f2(t0), z − f3(t0) f1(t0), f2(t0), f3(t0) f1 (t0), f2 (t0), f3 (t0) = 0 . 5.3 5.3. Now we can define the following objects at a non-inflection points p ∈ C: objekty: (i) The plane ν through the point p perpendicular to the tangent line is called normal plane. 28 (ii) The intersection n = ν ∩ ω of the normal and osculating planes is called principal normal line. (iii) The line b through the point p perpendicular to the osculating plane is called binormal line. (iv) The plane determined by the tangent and binormal lines is called rectifying plane. JS: missing picture de5.4 5.4 Definition. Non-inflection point p ∈ C is called planar point if the osculating plane at this point has the 3rd order contact with the curve C. ve5.5 5.5 Theorem. The non-inflection point f(t0) is planar if and only if the vector d3f(t0) dt3 is linearly independent on vectors df(t0) dt and d2f(t0) dt2 . Proof. Consider the function Φ(t) from the proof of the theorem 5.1 and two conditions in this proof for the first and second derivatives of Φ(t). Considering the 3rd order contact, we have moreover e5.1e5.1 (1) d3Φ(t0) dt3 = a d3f1(t0) dt3 + b d3f2(t0) dt3 + c d3f3(t0) dt3 = 0 . Thus the vector d3f(t0) dt3 lies in the associated vector space of the osculating plane therefore it is linearly dependent on vectors df(t0) dt and d2f(t0) dt2 . In the opposite direction, if this linear dependence holds then the equation (1) is a consequence of two equations from the proof of the theorem 5.1. Thus C has the 3rd order contact with the osculating plane. 5.6 5.6. Let us further consider the arc-length s. We have e1(s) = df(s) ds which is the unit vector. Considering a non-inflection point f(s), we denote by e2(s) the unit collinear with de1(s) ds in the same direction. Thus e5.2e5.2 (2) de1(s) ds = κ(s)e2(s) , κ(s) > 0 . and the vector e2(s) lies in the osculating plane. According to 1.26, the vector e2(s) is perpendicular to e1(s). Thus e2(s) yields the direction of the principal normal line at the point f(s). 5.7 5.7. Assume further the space E3 is oriented. By e3(s) we denote the unit vector perpendicular to e1(s) and e2(s) such that the basis e1(s), e2(s), e3(s) is positive. Thus the vector e3(s) yields the direction of the binormal. Definition. The frame f(s0), e1(s0), e2(s0), e3(s0) is called Frenet frame of the curve C in the non-inflection point f(s0). 29 The argument s will be further usually omitted. 5.8 5.8. Since the vector e2 is unit, by differentiating of the relation (e2, e2) = 1 we obtain e2, de2 ds = 0. Thus de2 ds = ce1 + τe3 . Similarly by differentiating (e1, e2) = 0 we get de1 ds , e2 + e1, de2 ds = 0 i.e. κ + c = 0. Thus e5.3e5.3 (3) de2(s) ds = −κ(s)e1(s) + τ(s)e3(s) . By differentiating (e3, e3) = 1 we see the vectpr de3 ds is perpendicular to e3. By differentiating of the relation (e1, e3) = 0 we obtain de1 ds , e3 + e1, de3 ds = 0 . But de1 ds = κe2 thus the first scalar product is zero, i.e. de3 ds = ke2. Finally by differentiating (e2, e3) = 0 we get de2 ds , e3 + e2, de3 ds = 0. From this it follows τ + k = 0 hence e5.4e5.4 (4) de3(s) ds = −τ(s)e2(s) . We have shown Theorem (Frenet formulae). The curve f(s) without inflection points satisfies e5.5e5.5 (5) df ds = e1, de1 ds = κe2, de2 ds = −κe1 +τe3, de3 ds = −τe2 . de5.9 5.9 Definition. The number κ(s0) > 0 is called curvature and the number τ(s0) is called torsion of the spacial curve f(s) in the non-inflection point f(s0). 5.10 5.10. Frenet formulae of the curve f(s) yield e5.6e5.6 (6) df ds = e1 , d2f ds2 = κe2 , d3f ds3 = dκ ds e2 + κ(−κe1 + τe3) . 30 Assume 0 ∈ I for the curve f(s), f : I → E3. Thus at s = 0 we have the Taylor expansion f(s) = f(0) + s e1(0) + κ(0)s2 2 e2(0) + s3 6 dκ(0) ds e2(0) − κ2 (0)e1(0) + κ(0) τ(0) e3(0) + ν(s) ,e5.7e5.7 (7) where ν(s) is the vector function whose value and first three derivatives are zero at the origin. In the other words, we have: Theorem. Let x, y, z be coordinates with respect to the Frenet frame f(0), e1(0), e2(0), e3(0) . Then the curve f(s) in a neighbourhood of the point f(0) is given by e5.8e5.8 (8) x = s − κ2(0) 6 s3 + ξ(s) , y = κ(0) 2 s2 + 1 6 dκ(0) ds s3 + η(s) , z = κ(0)τ(0) 6 s3 + ζ(s) , where real functions ξ(s), η(s) and ζ(s) have zero value and first three derivatives at the origin. Relations (8) yield so called local expansion of the curve f(s) with respect to its Frenet frame. Using this, we shall study orthogonal projections of the curve to its three basic planes of the Frenet frame. 5.11. A geometric meaning of the curvature of a plane curve is given by the definition 2.10. The curve C ≡ f(s) in E3 satisfies Theorem. Considering a non-inflection point p ∈ C, the curvature of the curve C is equal to the curvature of its orthogonal projections Cp to the osculating plane. Proof. We can assume p = f(0). Accoridng to (7), Cp has parametrization (or rather its Taylor expansion) e5.9e5.9 (9) x = s + α(s) , y = κ(0) 2 s2 + β(s) , where α(s) and β(s) have zero values and first two derivatives at the origin. The circle e5.10e5.10 (10) x2 + y − 1 κ(0) 2 = 1 κ(0) 2 , tj. x2 + y2 − 2 κ(0) y = 0 31 has the second order contact with Cp at the origin. Indeed, putting (9) JS: missing picture into (10) we get s2 − s2 + γ(s) = 0 , where γ(s) is a function which has zero value and first two derivatives at the origin. 5.12. Similarly, a parametrization of the orthogonal projection of the curve C to the normal plane os y = κ(0) 2 s2 + 1 6 dκ(0) ds s3 + η(s) , z = κ(0)τ(0) 6 s3 + ζ(s) . Denoting this vector valued function by g(s), we have dg(0) ds = o. Thus in a sense, the origin is an edge of the type of semicubic parabola. JS: missing picture 5.13. The orthogonal projection of C to its rectifying plane is x = s − κ2(0) 6 s3 + ξ(s) , z = κ(0)τ(0) 6 s3 + ζ(s) . Denoting by h(s) this vector valued function, we have dh(0) ds = (1, 0), d2h(0) ds2 = (0, 0). Hence the origin is an inflection point. JS: missing picture Comparision of all three projections yields a better understanding how the curve C moves “through” its Frenet frame. 5.14. It follows from Frenet formulae that planar points of the spacial curve C are easily characterized in terms of the torsion. Theorem. The point f(s0) is planar if and only if τ(s0) = 0. Proof. It follows from (6) that the vector d3f(s0) ds3 is a linear combination of e1(s0) a e2(s0) if and only if τ(s0) = 0. 5.15. The curve C ⊂ E3 is called planar if it lies in some plane ⊂ E3. Since C lies in lies, every point of C is planar hence the torsion of a plane curve is zero. The opposite direction follows from Frenet formulae. Theorem. A simple curve where all points are planar, is a part of a plane. 32 Proof. The condition τ = 0 yileds de3 ds = o i.e. e3 is a constant vector. Consider the plane through the point f(s0) perpendicular the vector e3(s0). Its equation is e3(s0), w − f(s0) = 0 where w = (x, y, z) is an arbitrary point in E3. Consider the function ϕ(s) = e3(s0), f(s) − f(s0) . We have dϕ ds = e3(s0), e1(s) = 0 since e3(s0) = e3(s). Thus ϕ is a constant function. Further ϕ(s0) = 0 thus thus the function ϕ is identically zero. The whole curve lies in the considered plane. 5.16. The basic geometrical meaning of the torsion follows firectly from (5). Theorem. It holds |τ| = de3 ds . Therefore we can say that torsion is the velocity of rotation of the binormal vector. Zero rotation of course indicates plane curves. GEnerally one can say that greater absolute value of torison, more the curve diverges from a plane curve. 5.17 5.17. We shall find a formula for the curvature κ with respect to an arbitrary parametrization f(t) of the curve C. It follows from (6) that κ = df ds × d2f ds2 . Put t = t(s). The chain rule yields e5.11e5.11 (11) df ds = df dt dt ds , d2f ds2 = d2f dt2 dt ds 2 + df dt d2t ds2 . Further we know dt ds = 1 df dt . Since the vector product of two colinear vectors is zero, we have df ds × d2f ds2 = df dt dt ds × d2f dt2 dt ds 2 + df dt d2t ds2 = df dt × d2f dt2 dt ds 3 . Thus we have proved Theorem. It holds e5.12e5.12 (12) κ = df dt × d2f dt2 df dt 3 5.18. We shall derive a formula for torsion τ with respect to an arbitrary parametrization f(t) of the curve C. Recall three vectors u = (u1, u2, u3), v = (v1, v2, v3), w = (w1, w2, w3) in the oriented three-dimensional Euclidean vector space determine the exterior product [u, v, w] = u1 u2 u3 v1 v2 v3 w1 w2 w3 . 33 Theorem. It holds e5.13e5.13 (13) τ = df dt , d2f dt2 , d3f dt3 df dt × d2f dt2 2 . Proof. First we observe the exterior product satisfies [u, v + au, w + bu + cv] = [u, v, w] . It follows from (6) that df ds , d2f ds2 , d3f ds3 = κ2 τ[e1, e2, e3] = κ2 τ , as e1, e2, e3 is a positive basis. We shall rewrite (11) (obtained in the proof of the theorem 5.17) in the form e5.14e5.14 (14) df ds = df dt dt ds , d2f ds2 = d2f dt2 dt ds 2 + g df dt , where g = d2t ds2 but we shall not need this fact. Further differentiation yields e5.15e5.15 (15) d3f ds3 = d3f dt3 dt ds 3 + h d2f dt2 + k df dt , where we shall not need coefficients h and k explicitly. Thus we have κ2 τ = df ds , d2f ds2 , d3f ds3 = dt ds df dt , dt ds 2 d2f dt2 , dt ds 3 d3f dt3 = = dt ds 6 df dt , d2f dt2 , d3f dt3 . Using (12) for κ and the relation dt ds = 1 df dt , (13) follows. 5.19 Example. We shall find curvature and torsion of the screw line. This curve is given by the trajectory of the uniform screw motion. Its JS: missing picture parametrization therefore is f(t) = (a cos t, a sin t, bt) , t ∈ (−∞, ∞), a > 0 . The number a is the radius of the circular cylinder on which the screw line lies. The number b is called slope of the screw line. 34 Consecutive differentiation yields f = (−a sin t, a cos t, b) , f = (−a cos t, −a sin t, 0) , f = (a sin t, −a cos t, 0) . Thus f × f = (ab sin t, −ab cos t, a2), f × f = a √ a2 + b2. Further f = √ a2 + b2. Podle (12) m me κ = a a2+b2 . To deteremine torison we compute the determinant [f , f , f ] = −a sin t a cos t b −a cos t −a sin t 0 a sin t −a cos t 0 = ba2 . According (13) we have τ = ba2 a2(a2+b2) = b a2+b2 . Thus the screw line has constant curvature and torsion. 5.20 5.20. Recall proper Euclidean transformation in an oriented space E3 is such Euclidean transformation ϕ: E3 → E3 which preserves orientation. Similarly as in a plane, we call two curves C, ¯C ⊂ E3 congruent if there exists such proper Euclidean transformation ϕ such that ϕ(C) = ¯C. As in the plane, we shall assume that C and ¯C are simple and both are parametried by the arc-length on the same interval I. Theorem. Let curves C and ¯C are without inflection points, f : I → E3 and ¯f : I → E3 are their arc-length parametrizationson the same interval I and κ(s), ¯κ(s) and τ(s), ¯τ(s) are tehir curactures and torsions, respectively. Then curves C and ¯C are conguent if and only if κ = ¯κ and τ = ¯τ on I. Proof. On one side, it directly follows from the geometric construction of the Frenet frame that given two congruent curves, their curvatures and torsions are the same functions of the arc-length. In the opposite direction, consider C and ¯C with Frenet frames f(s), e1(s), e2(s), e3(s) and ¯f(s), ¯e1(s), ¯e2(s), ¯e3(s) , respectively. Beside (5), we have also e5.16e5.16 (16) d ¯f ds = ¯e1 , d¯e1 ds = κ¯e2 , d¯e2 ds = −κ¯e1 + τ ¯e3 , d¯e3 ds = −τ ¯e2 with the same κ and τ. Thus (5) and (16) is the same system of differencial equations for twelve real functions which are components of f, e1, e2 and e3. Given s0 ∈ I, both f(s0), e1(s0), e2(s0), e3(s0) as well as ¯f(s0), ¯e1(s0), 35 ¯e2(s0), ¯e3(s0) are the point and a positive orthonormal frame. Hence there is a unique proper Euclidean motion ϕ: E3 → E3 which transforms the first of these 4-tuples to the second one. Then parametrization ¯f : I → E3 of the curve ¯C together with vector functions ¯e1, ¯e2 and ¯e3 and parametrizations ϕ ◦ f : I → E3 of the curve ϕ(C) together with vector functions ϕ ◦ e1, ϕ ◦ e2 and ϕ ◦ e3 satisfy the same system of differential equations with the same initial conditions. According to the theorem about unique existence of solution of a system of differential equations, we in particularly have ¯f = ϕ ◦ f. Thus ¯C = ϕ(C). 5.21 5.21. Similarly as in the plane we also have the opposite direction. Theorem. Let κ, τ : I → R be real functions, κ > 0. Then locally there exists a curve C parametrized by the arc-length on I such that κ is its curvature and τ is its torsion, 5.22 Example. We shall show the screw lines are unique curves with constant curvature and torsion (Zero torsion corresponds to a circle as the screw line with zero slope.) Indeed, we computed in (19) that e5.17e5.17 (17) κ = a a2 + b2 , τ = b a2 + b2 . for screw lines. Let κ > 0 and τ be given. Then we compute from (17) that τ κ = b a thus a = kκ, b = kτ for some k > 0. Putting this to the formula for κ w eget κ = kκ k2(κ2+τ2) , i.e. k = 1 κ2+τ2 . It follows from theorems 5.20 and 5.21 that parts of the screw lines with values a = κ κ2+τ2 and b = τ κ2+τ2 are unique curves with given constant κ and τ. 5.23 Remark. Another interesting (however less imporant in practice) geometrical object determined by the curve C ≡ f(s) is its osculating sphere. Assuming f(s0) is a non-planar point then there exists a unique sphere S which has the 3rd order contact with the curve C at f(s0). We shall only sketch its construction. It follows from the 1st order contact that the tangent line of the curve at the point f(s0) is also tangent to S hence the center of the osculating sphere lies on the normal line. Let f(s0) + ae2(s0) + be3(s0) is this center. We shall write the equation of the sphere S in the form of scalar product w − f(s0) − ae2(s0) − be3(s0), w − f(s0) − ae2(s0) + be3(s0) = a2 + b2 where w = (x, y, z) is an arbitrary point in E3. To determine the contact of C and S we use the function Φ(s) = f(s)−f(s0)−ae2(s0)−be3(s0), f(s)−f(s0)−ae2(s0)−be3(s0)−a2 −b2 . 36 Relations Φ(s0) = 0 and dΦ(s0) ds = 0 hold by construction. Conditions d2Φ(s0) ds2 = 0 and d3Φ(s0) ds3 = 0 yield e5.18e5.18 (18) a = 1 κ(s0) , b = − κ (s0) κ2(s0)τ(s0) , κ (s) = dκ ds . The radius r = √ a2 + b2 of the usculating sphere thus is e5.19e5.19 (19) r = 1 κ2|τ| κ2τ2 + dκ ds 2 . Display relations (18) and (19) illustrate an interesting general akt. According to theorems 5.20 and 5.21, the curve C is geometrically determined by its curvature and torsion. Thus also other geometrical objects deteremined by the curve and its invariants are expressed using κ and τ and their derivatives with respect to the arc-length. 37 6 The first fundamental form of the surface Now we start a systematic study of surfaces in E3. 6.1 6.1. Consider a surface S with a local parametric expression f(u1, u2), (u1, u2) ∈ D, see 4.4. We shall use the abbreviated notation f1 = ∂1f, f2 = ∂2f. Thus f1(u0), f2(u0) form a basis of the tangent space TpS of the surface S at the point p = f(u0). Consider vectors A, B ∈ TpS, A = a1f1 + a2f2, B = b1f1 + b2f2. Their scalar product is given by e6.1e6.1 (1) (A, B) = (a1f1 + a2f2, b1f1 + b2f2) . Put e6.2e6.2 (2) g11 = (f1, f1) , g12 = (f1, f2) , g22 = (f2, f2) . That is, gij, i, j = 1, 2 are functions on D. Then (1) can be written in the form e6.3e6.3 (3) (A, B) = g11a1b1 + g12(a1b2 + a2b1) + g22a2b2 . This is a bilinear form on TpS. The corresponding quadratic form deteremines the length of the vector A, A = g11a2 1 + 2g12a1a2 + g22a2 2 . The angle ϕ of vector A, B satisfies (4) cos ϕ = g11a1b1 + g12(a1b2 + a2b1) + g22a2b2 g11a2 1 + 2g12a1a2 + g22a2 2 g11b2 1 + 2g12b1b2 + g22b2 2 6.2. Consider a curve u(t) = u1(t), u2(t) on S. We have df dt = f1 du1 dt + f2 du2 dt hence df dt = g11 du1 dt 2 + 2g12 du1 dt du2 dt + g22 du2 dt 2 From the formula for length of an arc of a space curve we have Theorem. Length s of the arc of the curve u(t) on the surface f(u) between points with parameters t1 and t2 is (5) s = t2 t1 g11 du1 dt 2 + 2g12 du1 dt du2 dt + g22 du2 dt 2 dt . 38 Thus the differential ds is given by the expression which follows the symbolem of integral. Its square e6.6e6.6 (6) (ds)2 = g11(du1)2 + 2g12du1du2 + g22(du2)2 is a quadratic form corresponding to the bilinear form (3). 6.3 Definition. The quadratic form (6) is called first fundamental form of the suface. It is denoted by Φ1 or (ds)2. We shall use the same symbol Φ1 also for the bilinear form determined by this quadratic form. 6.4 6.4 Example. Consider the sphere S centered at the origin se with the radius r. Given a point p ∈ S away from the z-axis, consider its projection JS: missing picture q to the plane (x, y). We shall denote by u1 the angle of the radius vector of the point q with the positive half x-axis, i.e. u1 ∈ [0, 2π). We denote by u2 the angle of the radius vector of the point p with the plane (x, y), i.e. u2 ∈ − π 2 , π 2 . Thus z = r sin u2 and length of the radius vector of q is r cos u2. Situation in the plane (x, y) corresponds to polar coordinates, i.e. x = r cos u2 cos u1, y = r cos u2 sin u1. Summarizing, we have (7) f(u1, u2) = (r cos u1 cos u2, r sin u1 cos u2, r sin u2) , u1 ∈ (0, 2π), u2 ∈ − π 2 , π 2 . The sphere is a not a simple surface hence our parametrization does not cover half-circle which is intersection of the sphere with the half-plane x ≥ 0 in the (x, z)-plane. However, this incompletness is usually not a problem. We shall find the first fundamental form of the sphere. We have f1 = r(− sin u1 cos u2, cos u1 cos u2, 0) , f2 = r(− cos u1 sin u2, − sin u1 sin u2, cos u2) . Thus g11 = (f1, f1) = r2 cos2 u2, g12 = (f1, f2) = 0, g22 = r2. The first fundamental form of the sphere is of the form (8) Φ1 = r2 cos2 u2(du1)2 + (du2)2 . 6.5 6.5. We shall compute the first fundamental form of an explicitly given surface z = f(x, y), (x, y) ∈ D, see 4.5. Its parametrization is given by ¯f(x, y) = x, y, f(x, y) . Hence ¯f1 = (1, 0, fx), ¯f2 = (0, 1, fy) where fx and fy are partial derivatives of f with respect to x and y, respectively. Computing scalar products (2), we obtain (9) Φ1 = (1 + f2 y ) (dx)2 + 2fxfy dx dy + (1 + f2 y ) (dy)2 . 39 6.6 Definition. System of curves on a simple surface S is a 1-parameter family L of curves on S such that there is a unique curve from L through every point of the surface S. First consider system L in the parameter space D of the plane (u1, u2). Assume that tangent lines of curves of the system are not parallel with the u2-axis. Then system L(u1, u2) of tangent lines of the system L satisfy e6.10e6.10 (10) du2 du1 = L(u1, u2) . We say (10) is differential equation of the system L. Vector field on the space D is the rule which assignes a vector in the tangent space TpD for every point p ∈ D. Having a nowhere vanishing vector field F1(u1, u2), F2(u1, u2) on D tangent to the system L, non-parllelity with the u2-axis mean F1(u1, u2) = 0. Then the differential equation of the system L is (11) du2 du1 = F2(u1, u2) F1(u1, u2) . The system L on a surface S is usually given on the parameter space. Vector field on the surface S is a rule which assignes a vector in the tangent space TpS to every point p ∈ S. 6.7 Definition. Orthogonal trajectories of the system L on the surface S is a system L ’ na S such that curves of L and L ’ are perpendicular at every point. Theorem. If F1(u1, u2), F2(u,u2) is a coordinate expression of a vector field tangent to the system L then the differential system of its orthogonal trajectories is e6.12e6.12 (12) du2 du1 = − g11F1 + g12F2 g12F1 + g22F2 . Proof. Let (du1, du2) is a tangent vector to the required system L ’. Following (3), the orthogonality of both systems means g11F1du1 + g12(F1du2 + F2du1) + g22F2du2 = 0 . Now (12) follows by an algebraic manipulation. 40 Remark. If there zero in the denominator of the right hand side of (12) at some point of the surface, it means orthogonal trajectories through this point are, considered at the parameter space, tangent to the axis u2. Then we should consider the differential equation of the system with interchanged axes u1 and u2. 6.8 Example. We shall find orthogonal trajectories for the system u1 + u2 = konst on the sphere from 6.4. By differentiation be find du1 +du2 = 0, i.e. the differential equation of this system is du2 du1 = −1. Thus we cna put F1 = 1 and F2 = −1. We found g11 = r2 cos2 u2, g12 = 0, g22 = r2 an 6.4. Following (2), the differential equation of orthogonal trajectories is e6.13e6.13 (13) du2 du1 = cos2 u2 . Separating variables in (13) and integrating, we obtain the equation of orthogonal trajectories in the form tg u2 = u1 + konst. 6.9 Definition. Net on the surface S are two systems L1, L2 whose curves have nonzero angle at every point. The net is called orthogonal if this is the right angle at every point. A simple example is the parametric coordinate net formed by curves u1 = konst. and u2 = konst. given by the parametrization f(u1, u2) of the surface S. The condition f1 ×f2 = o guarantees that angle of curves of two parametric systems is nonzero. The following statement will be useful in many specific case. Theorem. The parametric net is orthogonal if and only if g12 = 0. Proof. Vectors f1 and f2 are tangent to parametric systems and g12 = (f1, f2). 6.10 6.10 Lemma. It holds g11g22 − g2 12 > 0. Proof. The Cauchy inequality says that vectors a, b satisfy |(a, b)| ≤ a b , i.e. (a, b)2 ≤ a 2 b 2. Here the equality holds only if these vectors are collinear. Putting a = f1, b = f2, we have a 2 = g11, b 2 = g22, (a, b) = g12. Since vectors f1 and f2 are not collinear, the lemma fol- lows. 41 6.11. A standard result in calculus shows that the area of the surface given explicitly as z = f(x, y), (x, y) ∈ D (where f is a bounded function on a bounded space D) is given by the double integral e6.14e6.14 (14) D 1 + f2 x + f2 y dx dy 6.12. The map f : D → E3 is bounded if the set f(D) lies in some ball. Theorem. Let the surface S is given by a bounded map f(u1, u2) on a bounded space D ⊂ R2. Then its area is given by e6.15e6.15 (15) D g11g22 − g2 12 du1 du2 . Proof. We know every surface cen be given explicitly in a neighbourhood of every point and let z = f(x, y) be such parametrization. We have shown g11 = 1+f2 x, g12 = fxfy, g22 = 1+fy2 in 6.5, hence g11g22−g2 12 = 1+f2 x +f2 y . Then (15) localy reduces to the usual expresssion redukuje to the usual expression (14). The global version follows from the additivity of area of the surface. The expression dV : = g11g22 − g2 12 du1 du2 is also called volume element of the surface S. The formula of the area of the surface thus has the form V = D dV. 6.13 Example. We shall find area V of the so called spherical cap on the sursface with the radius r with the angle α, see the picture. Thus JS: missing picture D = (0, 2π) × π 2 − α, π 2 . We found g11 = r2 cos2 u2, g12 = r2 in 6.4. Thus V = D r2 cos u2 du1 du2 = r2 2π 0 du1 π/2 π/2 − α cos u2du2 = 2πr2 sin u2 π/2 π/2 − α = 2πr2 (1 − cos α) . The case α = π 2 gives the area 2πr2 of the half of the sphere. 6.14. Summarizing, the first fundamental form Φ1, determined by scalar products at each tangent space of the surface, is used mainly for computation of length of curves on surfaces, angle of these curves and area of the surface. A fundamental theoretical meaning of Φ1 will be discussed later. 42 7 The second fundamental form of the surface 2fundamental 7.1 7.1. Consuider the normal line NpS of the surface S at the point p. We have two unit vectors on the normal line, a choice of one of them yields orientation of the line NpS. Definition. Orientation of the surface S is a choice of orientation of every normal line in a continuous way. One can orient every simple surface. Given a parametrization f(u1, u2), we can chose the direction of the normal as the direction of the vector product f1 × f2. 7.2. The case of M¨obius strip shows that there surface which cannot be oriented. Definition. The surface S, which can be oriented, is called orientable. An orientable surface together with a choice of orientation is called oriented. We shall denote the unit vector of the oriented normal line by n. Dependence on parameters is expressed if we write n(u1, u2). In the case of orientation determined by the parametrization f(u), we have e7.1e7.1 (1) n = f1 × f2 f1 × f2 . The orthogonality of the normal line and the surface is expressed by equa- tions e7.2e7.2 (2) (n, f1) = 0 , (n, f2) = 0 . 7.3 7.3. Further we assume the surface S is oriented. Given an arbitrary motion γ(t) in the space E3, the vector d2γ dt2 is called accelaration. Consider a motion on the surafce S with a local parametrizaiton f(u) given by u1(t), u2(t) in the parameter space D. This is the motion γ(t) = f u1(t), u2(t) in E3. Let us compute teh acceleration. The first derivative of the composed function is given by the expression dγ dt = f1 u1(t), u2(t) du1 dt + f2 u1(t), u2(t) du2 dt . We shall use an abbreviation e7.3e7.3 (3) f11 = ∂11f , f12 = ∂12f , f22 = ∂22f . 43 to compute the second derivative. We obtain e7.4e7.4 (4) d2γ dt2 = f11 du1 dt 2 + 2f12 du1 dt du2 dt + f22 du2 dt 2 + f1 d2u1 dt2 + f2 d2u2 dt2 . Following (2), the scalar product of n and d2γ dt2 depends only on dγ dt . Definition. The scalar product n, d2γ dt2 is called normal acceleration corresponding to the vector dγ dt ∈ TpS. In the case of dγ dt =1, this is termed normal curvature of the oriented surface S in the direction of this vector. That is, the sign of the normal acceleration depends on orientation fo the surface. 7.4 7.4. Consider scalar products e7.5e7.5 (5) h11 = (n, f11) , h12 = (n, f12) , h22 = (n, f22) , which are functions on the space D. Following (4), we obtain the rule which which associates the normal acceleration to every tangent vector (du1, du2) ∈ TpS. This is the quadratic form on TpS on the space e7.6e7.6 (6) h11(du1)2 + 2h12du1 du2 + h22(du2)2 . Definition. The quadratic form (6) is called second fundamental form of the surface S and will be denoted by Φ2. That is, the second fundamental form of the oriented surface S is a rule which maps every vetor A ∈ TpS to the number Φ2(A) which he obtained as follows. We consider a motion γ(t) on the surface S such that A = dγ(t0) dt . We compute its acceleration d2γ(t0) dt2 . The number Φ2(A) is then equal to the scalar product n(γ(t0)), d2γ(t0) dt2 where n(γ(t0)) is the oriented vector of the normal line at the point γ(t0). 7.5 7.5. Consider the direction of the nonzero vector A in the tangent space TpS. The section of the surface S by the plane determined by the normal line NpS and the direction A is the curve which we call normal section of the surface in direction A. The basic geometrical meaningof the form Φ2 is given by Theorem. The absolute value of the normal curvature in the direction of the vector A is equal to the curvature of the normal section in this direction. 44 Proof. Consider the parametrization γ(s) of this section by the arc-length, γ(s0) = p. Then dγ ds is a unit vector and d2γ(s0) ds2 is perpendicular to this vector. We know (form theory of curves) that the norm of d2γ(s0) ds2 is equal to the curvature of the normal section. Vectors n(p) and d2γ(s0) ds2 are thus colinear. Since n(p) is a unit vector, the absolute value of the scalar product n(p), d2γ(s0) ds2 is equal to the norm of the second vector. 7.6. The normal curvature κ in the direction of the vector A = (du1, du2) satisfies e7.7e7.7 (7) κ = h11(du1)2 + 2h12du1 du2 + h22(du2)2 g11(du1)2 + 2g12du1 du2 + g22(du2)2 . Indeed, the unit vector in this direction is 1 A (du1, du2) where A 2 = g11(du1)2 + 2g12du1 du2 + g22(du2)2. Substituting this into (6), the display (7) follows. 7.7 Definition. The point f(u0) ∈ S is called planar point if the form Φ2(u0) is nonzero, i.e. h11(u0) = 0, h12(u0) = 0, h22(u0) = 0. 7.8 Definition. The surface S is called connected if every two its point can be connected by a motion which lies in S. 7.9 7.9 Theorem. A simple connected surface S with all points planar is a part of a plane. Proof. Pur n1 = ∂1n, n2 = ∂2n. Differentiating (2) with respect to u1 and u2 yields e7.8e7.8 (8) (n1, f1) + (n, f11) = 0 , (n1, f2) + (n, f12) = 0 , (n2, f1) + (n, f12) = 0 , (n2, f2) + (n, f22) = 0 . In particular, this shows that all points on a plane (with constant normal vector) are planar. Further we use the fact that n is a unit vector. By differentiating the relation (n, n) = 1, we obtain e7.9e7.9 (9) (n, n1) = 0 , (n, n2) = 0 . If every point of the surface S is planar, the second term in every relation of (8) is zero, cf. (5). Then first two equations of (8) and the first equation in (9) mean the vector n1 is perpendicular to three lienarly independent vectors n, f1, f2. Thus n1 is the zero vector. Analogously, it follows from 45 remaining equations of (8) and (9) that n2 is the zero vector. Thus the normal vector is constant, n = a. Consider the function ϕ(u1, u2) = a, f(u1, u2) − f(u0 1, u0 2) . We have ∂ϕ ∂u1 = (a, f1) = 0, ∂ϕ ∂u2 (a, f2) = 0 hence ϕ is a constant function. Moreover, ϕ(u0 1, u0 2) = 0, i.e. ϕ(u) = 0 for all u. This means that the whole surface S lies on its tangent plane through the point f(u0). 7.10 Definition. The point f(u0) ∈ S is called spherical point if the form Φ2(u0) is a constant multiple of the form Φ1(u0). Thus the spherical point f(u0) is characterized by the condition e7.10e7.10 (10) h11(u0) = cg11(u0) , h12(u0) = cg12(u0) , h22(u0) = cg22(u0) , 0 = c ∈ R . The normal vector n(u) of the sphere centered at the origin with raidus r satisfies n(u) = 1 r f(u). Equations (8) show that all points on the sphere are spherical. 7.11 Theorem. A simple connected surface S where all points are spherical, is a part of a sphere. Proof. Assume (10) holds. That is, e7.11e7.11 (11) (n, f11) = c(f1, f1) , (n, f12) = c(f1, f2) , (n, f22) = c(f2, f2) . It follows from (8) and (11) thast e7.12e7.12 (12) (f1, n1+cf1) = 0 , (f1, n2+cf1) = 0 , (f2, n1+cf2) = 0 , (f2, n2+cf2) = 0 . Further, it follows from (2) and (9) that e7.13e7.13 (13) (n, n1 + cf1) = 0 , (n, n2 + cf2) = 0 . Analogously as in the proof of theorem 7.9, it follows from these relations that e7.14e7.14 (14) n1 + cf1 = o , n2 + cf2 = o . By differentiation of the first equation with respect to u2 and the second equation with resoect to u1, we obtain (15) n12 + ∂c ∂u2 f1 + cf12 = 0 , n12 + ∂c ∂u1 f2 + cf12 = 0 , 46 where n12 = ∂2n ∂u1∂u2 . Thus we have the difference ∂c ∂u2 f1 − ∂c ∂u1 f2 = o . Since vectors f1 and f2 are linearly independent, we have ∂c ∂u1 = 0, ∂c ∂u2 = 0, i.e. c is a constant. Following (14), the point f + 1 c n is fixed. The distance of every point on the surface form this point is constant and equal to 1 |c| , i.e. S is a part of the corresponding sphere. de7.12 7.12 Definition. A direction in the tangent plane of the surafce is called asymptotic direction, if its normal curvature is zero. The tangent line in this direction is called asymptotic tangent line. Thus the equation of asymptotic directions is e7.16e7.16 (16) h11(du1)2 + 2h12du1 du2 + h22(du2)2 = 0 Every direction is asymptoic in planar points. Assuming the direction du2 = 0 is not asymptotic, i.e. h11 = 0, put = du1 du2 . Then (16) yields the quadratic equation for asymptotic directions e17e17 (17) h11 2 + 2h12 + h22 = 0 . Its roots satisfy 1,2 = −h12± √ h2 12−h11h22 h11 . Put e7.18e7.18 (18) h = h11 h12 h12 h22 = h11h22 − h2 12 . Thus there are two (real) asymptotic directions for h < 0, both directions coincide for h = 0 and there are imaginary roots for h > 0. If h11 = 0 and h22 = 0, (16) yields the quadratic equation for the fraction du2 du1 and the situation is similar. If h11 = 0 and h22 = 0, we have h12 = 0 in a non-planar point, i.e. asymptotic directions are du1 = 0 and du2 = 0. de7.13 7.13 Definition. The non-planar point is called hyperbolic or parabolic or elliptic, if h < 0 or h = 0 or h > 0, respectively. In spherical points, the inequality 6.10 means h > 0 hence this is a special case of an elliptic point. 7.14 Definition. The curve C on the surface S is called asymptotic if its tangent line ate every point is asymptotic tangent line. 47 Thus we have two systems of asymptotic curves on surfaces with only hyperbolic points, one system of asymptotic curves on surfaces with only parabolic points and there is no system of asymptotic curves on surfaces with only elliptic points. 7.15 Theorem. A line in the tangent plane τpS is an asymptoticc line if and only if it has contact of the second order with the surface. Proof. If a direction is asymptotic then the normal section in this direction has zero curvature at the point p. Thus p is inflection point of the normal section, i.e. the its tangent line has contact of the second order with this section. In the opposite direction, if a tangent line at the point p ∈ S has contact of the 2nd order with some curve γ(t) on S, γ(t0) = p, it is the inflection point of this curve. Thus the vector d2γ(t) dt2 is colinear with vector dγ(t0) dt which is perpendicular to the normal vector n(p), i.e. e7.19e7.19 (19) n(p), d2γ(t0) dt2 = 0 . 7.16. Recall the osculating plane of a spacial curve is not determined in its inflection points, Theorem. A curve C on the surface S is asymptotic if and only if, at each its point, the osculating plane coincides with the tangent plane of the surface or is not determined. Proof. The osculating plane of the curve C ≡ γ(t) at the point p = γ(t0) is determined by vectors dγ(t0) dt , d2γ(t0) dt2 if these are linearly independent. Here dγ(t0) dt llies in the tangent plane of the surface. Thus the tangent plane of S coincides with the osculating plane of the curve C if and only if teh normal vector n(s) is perpendicular to d2γ(t0) dt2 , i.e. (19) holds. If this is an inflection point, the vector d2γ(t0) dt2 is colinear with dγ(t0) dt and (19) holds as well. In the opposite direction, if Φ2 dγ(t0) dt = 0 then (19) holds and similarly as in the first part of the proof, one verifies this is one of two cases obtained above. 7.17. We know from 1.28 that a (part of a) line is characterized by the fact that all its poins are inflection. Thus if there is a (part of a) line on the surface, it is an asymptotic curve. This yields e.g. asymptotic directions and curves on regular ruled surfaces, i.e. on the hyperboloid of one sheet and hyperbolic paraboloid. 48 7.18. Considering hyperbolic points, asymptotic directions divide directions in the tangent plane to two parts The sign of the normal curvature is JS: missing picture positive in one of them and negative in the other one. Thus in the positive part, normal sections lie locally above the tangent plane in the direction of the oriented normal line and in the negative part, normal section lie on the opposite part of the tangent plane. Thus the surface lies on both sides of the tangent plane. A prominent example is the surface z = xy. Axes x and y lie on this surface, i.e. they are asymptotic curves. The tangent plane at the origin is z = 0. Assuming x > 0, y > 0 or x < 0, y < 0, the surface lies above the tangent plane, assuming x > 0, y < 0 nebo x < 0, y > 0, the surface lies under the tangent plane. The sign of the curvature in an elliptic point is the same in all directions hence the whole surface locally lies on one side of the tangent plane. The simples cases are sphere and ellipsoid. Another interesting example is the anuloid. On the “out part of the tire”, the surface lies on one side of the tangent plane, i.e. all points are elliptic there. The the whole inner part of the anuloid lies locally lies on both sides of the tangent plane, i.e. these are hyperbolic points. “Bottom and top” circles are formed by parabolic points. 7.19 Remark. Finally we shall show how one can characterize planar and spherical point using the notionof contact of surfaces. Let p be a common point of S and ¯S. We say surfaces S and ¯S have contact of the order k at the point p if for every curve C ⊂ S through the point p there exists a curve ¯C ⊂ ¯S such that curves C and ¯C contact of the kth order at the point p. It is shown in [5] that this is an equivalence relation and also a computational criterion (similar to the case of curves and surfaces in 2.5 and 4.7) is derived. Assuming the surface S is given by a parametrization f(u) and the surface ¯S is given by an equation F(x, y, z) = 0 then we shall form a function of two variables Φ(u1, u2) = F f1(u1, u2), f2(u1, u2), f3(u1, u2) . It holds that if surfaces S and ¯S have contact of the kth order at the joint point p = f(u0) if and only if all partial derivatives of the function Φ at the point u0 = (u0 1, u0 2) up to the order ≤ k are zero. The case k = 1 then means two surfaces have contact of the 1st order at a joint pont if and only if they have the same tangent line at this point. 49 Considering the surface ¯S, ax + by + cz + d = 0 , we have Φ(u1, u2) = af1(u1, u2), bf2(u1, u2), cf3(u1, u2) + d . Conditions for the contact of the first order a∂1f1(u0)+b∂1f2(u0)+c∂1f3(u0) = 0 , a∂2f1(u0)+b∂2f2(u0)+c∂2f3(u0) = 0 mean that the vector (a, b, c) is colinear with the normal vectorem n(u0) of the surface S at the point f(u0). The condition for contact of the 2nd order is n(u0), ∂11f(u0) = 0 , n(u0), ∂12f(u0) = 0 , n(u0), ∂22f(u0) = 0 . Thus the point p ∈ S is planar if and only if the tangent plane at this point has contact of the 2nd with the surface. A similar computation shows that the point f(u0) ∈ S is spherical if and only if there exists a sphere Q such that S and Q have contact of the 2nd order at the point f(u0). 50 8 Principal curves principal 8.1. Values of the normal curvature of the surface S ≡ f(u) at its nonplanar point p can be visualize in the folloing way. We consider the segment of the length 1√ |κ| on the tangent line (in both directions) in a nonasymptotical direction. Here 1√ |κ| denotes the normal curvature in this direction. If a1f1(p) + a2f2(p) is such a vector, square of its norm is 1 |κ| , i.e. e8.1e8.1 (1) g11a2 1 + 2g12a1a2 + g22a2 2 = 1 |κ| . But κ is given by 7.(7) hence (1) is equivalent to rovnici e8.2e8.2 (2) |h11a2 1 + 2h12a1a2 + h22a2 2| = 1 . Definition. The curve (2) is called Dupin indikatrix at a non-planar point of the surface. 8.2 8.2. The curve (2) is an ellipse in elliptic points. Considering the equation of the unit circle in our affine coordinates in our tangent plane is g11a2 1 + 2g12a1a2 + g22a2 2 = 1 then it follows from 7.(10) that the ellipse is a circle precisely in spherical points of the surface. Considering a hyperbolic point and changing suitable coordinates, we can transform the equation h11a2 1 + 2h12a1a2 + h22a2 2 = 1 to the form e8.3e8.3 (3) x2 a2 − y2 b2 = 1 . Thus (2) corresponds to a pair of so called conjugated hyperbolas which JS: missing picture is formed by (3) and the hyperbola x2 a2 − z2 b2 = −1. Considering a parabolic point, (2) is a pair of parallel lines in the tangent plane with the point of the surface in the middle. Indeed, we have h11h22 = h2 12 in this case. Considering h11 > 0, h12 > 0, we have h12 = ± √ h11 √ h22. Consider the case of the positive sign first. Then the equation (2) has the form e8.4e8.4 (4) 1 = h11a2 1 + 2 h11 h22a1a2 + h22a2 2 = h11a1 + h22a2 2 . This is an equation of the pair of parallel lines (5) 1 = h11a1 + h22a2 , −1 = h11a1 + h22a2 51 with the origin in the middle. The case of the negative sign yields the sema result. A similar computation yields the same result for h11 < 0, h22 < 0. 8.3. Assuming a non-spherical point, we define axes of the Dupin indikatrix as axes of the ellipse or common axes of the pair of conjugated JS: missing picture hyperbolas or as the axis of the pair of parallel lines together with the parpendicular line through the origin. Definition. Directions of the Dupin indikatrix are called principal directions of the surface S at a given point. A curve on S which touches a principal direction at every point is called principal curve. Principal directions are not defined in planar and spherical points. We thus have the net of principal curves on surfaces without planar and spherical points. This net is orthogonal. 8.4 8.4. Since Φ2 is a quadratic form, it deteremines the polar bilinear form denoted by the same symbol. Given two vectors A = (a1, a2), B = (b1, b2) ∈ TpS, we have e8.6e8.6 (6) Φ2(A, B) = h11(p)a1b1 + h12(p)(a1b2 + a2b1) + h22(p)a2b2 . The condition Φ2(A, B) = 0 depends only on directions of the vectors A, B. This is the condition of polar conjugation with respect to Φ2(p). Definition. Directions in the tangent plane determined by nonzero vectors A, B ∈ TpS are called conjugated if they are polar conjugated with respect to Φ2(p). From the computational point of view, the condition of conjugation is given by (6) beeing zero. ve8.5 8.5 Theorem. Principal directions of the surface are directions which are in the same time conjugated and perpendicular. Proof. We know from analytic geometry that this characterizes axes ellipses and hyperbolas. The case of parallel lines can be computed analogously. 8.6 8.6. Beside (6) beeing zero, principal directions satisfy also the condition of orthogonality e8.7e8.7 (7) Φ1(A, B) = g11a1b1 + g12(a1b2 + a2b1) + g22a2b2 = 0 . 52 If (b1, b2) is a nonzero direction satisfying (7) and for which (6) is zero, we have a system of two homogeneous linear equations with a nonzero solution. The determinant of the system is zero, e8.8e8.8 (8) g11a1 + g12a2, g12a1 + g22a2 h11a1 + h12a2, h12a1 + h22a2 = 0 . Paasing to the diferentials du1 = a1, du2 = a2, we get Theorem. The differential equation of the net of principal curves is e8.9e8.9 (9) g11du1 + g12du2, g12du1 + g22du2 h11du1 + h12du2, h12du1 + h22du2 = 0 . Since (9) is generally a quadratic equation for th fraction du2 du1 . Its two solutions du2 du1 = F1(u1, u2), du2 du1 = F2(u1, u2) are diferential equations of these two systems of principal curves. de8.7 8.7 Definition. Normal curvatures κ1, κ2 in principal directions are called principal curvatures of the surface. The sum H = κ1 +κ2 of principal curvatures is called mean curvature, the product K = κ1κ2 is called Gauss (or total) curvature. Considering spherical points, the normal curvature has the same value κ. Here we define H = 2κ, K = κ2. Considering planar points, all normal curvatures are zero. Here we put H = 0, K = 0. A change of orientation of the surface S changes the sign of the normal curvature. The sign of the mean curvature H thus depeneds on the orientation of the surface, however the sign of the Gauss curvature K is independent on the orientation of the surface. 8.8 8.8. The Dupin indikatrix shows that the normal curvature has extremals at principal directions. We shall use that to derive a formula for principal curvatures. The following computation concerns only the “generic” case but one can show that the result holds in all cases. Considering the direction = du1 du2 then according to 7.(7), the normal curvature κ( ) in this direction satisfies κ( ) = h11 2 + 2h12 + h22 g11 2 + 2g12 + g22 . To simplify the computation, we shall write this in the form e8.10e8.10 (10) κ(g11 2 + 2g12 + g22) − (h11 2 + 2h12 + h22) = 0 . 53 Differentiating this with respect to and using the condition fors extremals dκ d = 0, we get e8.11e8.11 (11) κ(g11 + g12) − (h11 + h12 = 0 . Multiplying this by to − and adding the result to (10), we odtain e8.12e8.12 (12) κ(g12 + g22) − (h12 + h22) = 0 Putting back = du1 du2 and using an algebraic manipulation, (11) and (12) have the form e8.13e8.13 (13) (κg11 − h11) du1 + (κg12 − h12) du2 = 0 (κg12 − h12) du1 + (κg22 − h22) du2 = 0 . Here (du1, du2) is a nonzero direction which realizes the extremal. Thus the determinant of the system of two linear equations (13) must be zero. Therefore Theorem. Principal curvatures κ1, κ2 are roots of the quadratic equation e8.14e8.14 (14) κg11 − h11, κg12 − h12 κg12 − h12, κg22 − h22 = 0 . 8.9 8.9. A simple corollary of (14) is Theorem. The mean and the Gauss curvature satisfy e8.15e8.15 (15) H = g11h22 − 2g12h12 + g22h11 g11g22 − g2 12 , K = h11h22 − h2 12 g11g22 − g2 12 . Proof. It follows from (14) that κ2 (g11g22 − g2 12) − κ(g11h22 − 2g12h12 + g22h11) + (h11h22 − h2 12) = 0 . The sum H = κ1 + κ2 and the product K = κ1κ2 of roots have the form (15) using a well known properties of of quadratic equations. kvadratick rovnice. Further we shall show that (15) holds also in spherical and planar points. According to 7.(7), a spherical point satisfies hij = κgij, i = 1, 2 where κ is a common value of the normal curvature in all directions. Then (15) means H = 2κ, K = κ2. We have hij = 0 in planar points, i.e. H = 0 and K = 0. 54 8.10 8.10. Since g11g22 − g2 12 > 0 and h11h22 − h2 12 is the expression used in the definition 7.13, we obtained a new insight in this definition. Corollary. Elliptic, parabolic and hyperbolic pointsare characterized by the condition K > 0, K = 0 and K < 0, respectively. Remark. Since K = 0 in planar points, planar points are sometimes considered as parabolic points. pr8.11 8.11 Example. The Gauss curvatureof the sphere with radius r is 1 r2 . Indeed, all its points are spherical and the normal section in every direction is a circle with radius r. Thus K = 1 r2 . 8.12 8.12. The following formula nicely describes the normal curvature in an arbitrary direction using principal curvatures. Theorem (Euler formula). Let σ1 and σ2 are principal directions at the point p of the surface S, let κ1 and κ2 are corresponding principal curvatures and let s be the direction which has the angle σ1 with ϕ. Then the normal curvature κs in this direction satisfies e8.16e8.16 (16) κs = κ1 cos2 ϕ + κ2 sin2 ϕ . Proof. Let e1, e2 be unit vectors in directions σ1, σ2, respectively. We can consider parameters u1, u2 on S such that e1 and e2 are tangent vectors of the parametric net, i.e. e1 = (du1, 0), e2 = (0, du2). Then g11(p) = g22(p) = 1, g12(p) = 0 and conjugacy of directions σ1 and σ2 yields h12(p) = 0. It follows from the general formula 7.(7) for κ that κ1 = h11(p), κ2 = h22(p). The unit vector in the direction s has the form e1 cos ϕ + e2 sin ϕ. Putting this vector into 7.(7), we get κs = κ1 cos2 ϕ + κ2 sin2 ϕ. 8.13 8.13. We shall discuss one more geometric property which directly characterizes principal curves. Given a curve γ(t) on the surface S, we denote by nγ the 1-parameter system of normal vectors along γ. Theorem. The curve γ(t) is a principal curve of the surface S if and only if the vector dnγ dt is colinear with the vector dγ dt for all t. Proof. Let S be given by the parametrization f(u) and γ is given by u1(t), u2(t) in the parameter space. That is, e8.17e8.17 (17) dγ dt = f1 du1 dt + f2 du2 dt . 55 Assume the vector e8.18e8.18 (18) a1f1 + a2f2 is perpendicular to (17). Similarly we have nγ(t) = n u1(t), u2(t) hence e8.19e8.19 (19) dnγ dt = n1 du1 dt + n2 du2 dt . This vector lies in the tangent plane since vectors n1 and n2 are perpendicular to n, see 7.(9). Vectors (17) and (19) are colinear if and only if vectors (18) and (19) are perpendicular. Using the formula 7.(8), we get 0 = f1a1 + f2a2, n1 du1 dt + n2 du2 dt = − h11a1 du1 dt + h12 a2 du1 dt + a1 du2 dt + h22a2 du2 dt .e8.20e8.20 (20) Thus directions du1 dt , du2 dt and (a1, a2) are orthogonal and conjugate, i.e. they are principal directions. Thus γ(t) is a principal curve. In the opposite direction, if γ(t) is a principal curve then the vector dγ dt will be conjugated with the perpendicular vector, i.e. the second equation in (20) holds. Then the first equation (20) implies the vector dnγ dt is colinear with the vector dγ dt for all t. 8.14 8.14 Example. Consider a surface of revolution S given by rotation JS: missing picture of a planar curve C around an axis in the same plane and does not intersect the curve. Similarly as in the earth, circles of latitude on S are formed by rotation of a particular point of the curve C, wheres meridians on JS: missing picture S are positions of the curve C in particular moments of the rotation. We shall show that circles of latitude and meridians are principal curves of the surface of revolution S. Consider an arbitrary meridian of the surface S which we identify with the curve C. Thus normal vectors nC(t) of the plane curve C are in the same time normal vectors of the surface. All vectors nC(t) are unit thus nC(t), nC(t) = 1 shows, by differentiation, that vectors dnC (t) dt are perpendicular to nC(t). Thus vectors dnC (t) dt are colinear with tangent vectors of the curve C. Thus meridians are principal curves according to the theorem 8.13. Circles of latitude are perpendicualr to meridians hence theu are also principal curves (because the net of principal curves is orthogonal). 8.15 8.15. One can derive the previous result also by computation. We prescribe the curve C in the plane (x, z) locally by the parametrization x = g(t), 56 z = h(t), t ∈ I, i.e. the two dimensional vector g (t), h (t) is nonzero for JS: missing picture each t ∈ I. We can assume values of the parameter t are positive. We shall rotate along the z-axis and the assumption that C does not intersect the axis of rotation, means that C lies in the half-plane x > 0, i.e. g(t) > 0 for all t ∈ I. We denote by v the angle between the projection of the rotation plane to the (x, y)-plane with the positive x-half-axis. The parameter space D can be viewed as the space between two circles in R2 which, in polar coordinates, is characterized by the length of the radius in the intervalu I with arbitrary polar angle. In this sense, we can write v ∈ [0, 2π). The point with the x-coordinate g(t) moves along the circle x = g(t) cos v, y = g(t) sin v in the plane z = h(t). Thus the parametric description of the surface of revolution is f(t, v) = g(t) cos v, g(t) sin v, h(t) , t ∈ I, v ∈ [0, 2π) . From the point if view of the general theory, t and v play the role of parameters u1 and u2, respectively. Partial derivatives with respect to t and v are f1 = (g cos v, g sin v, h ) , f2 = g(− sin v, cos v, 0) . Coefficients of the first fundamental form hence are g11 = g 2 + h 2 , g12 = 0 , g22 = g2 . Further we have f1×f2 = g(−h cos v, −h sin v, g ) , n = 1 g 2 + h 2 (−h cos v, −h sin v, g ) . In the second order we have partial derivatives f11 = (g cos v, g sin v, h ) f12 = g (− sin v, cos v, 0) , f22 = g(− cos v, − sin v, 0) . Following 7. (5), coefficients of the second fundamenta form are h11 = g h − h g g 2 + h 2 , h12 = 0 , h22 = h g g 2 + h 2 . 57 Generally, already conditions g12 = 0, h12 = 0 simplify the differential equation of principal curves (9) to the form g11 g12 h11 h22 du1 du2 = 0 . The determinant is zero if and only if h11 = cg11, h22 = cg22, i.e. this is a spherical or planar point which is excluded from consideration leading to spherical curves. The equation du1du2 = 0 then characterizes the parametric net u1 = konst. and u2 = konst. In our case of the surface of revoution, these are meridians and circles of revolution. 8.16 8.16. We shall describe a relation between the curvature of an arbitrary plane section of the surface S and the curvature of the normal section in the same direction. Let be an arbitrary plane through the point p ∈ S different form the tangent plane τpS. Theorem (Meusnierova). Let κn be the normal curvature of the surface S in the direction of the line ∩ τpS and 0 ≤ α < π 2 be the angle between the normal line NpS and the plane . Then the curvature κ of the section of the surface S by the plane at the point p satisfies κn = κ cos α . Proof. Let γ(s) be the arc-length parametrization of the curve ∩S, γ(0) = p. Following the theorem 7.5, we have κn = n, d2γ(0) ds2 . It follows from the theory of planar curves that d2γ(0) ds2 = κ e2 where e2 is a unit vector in the plane . in our case, we have |(n, e2)| = cos α. 8.17 8.17. Consider a non-asymptotic direction A in the tangent plane. Denote by cn the center of curvature of the normal section and by c the center of curvature of the sectionby the plane , see the picture which shows JS: missing picture the section by the plane perpendicular to teh direction A. It follows form the Meusnier therem that cos α = κ κ , hence the triangle p cn c has the right angle at the vertex c . Geometrically this means that centers of the curvature of all plane sections of the surface S in the direction A lie on a circle for which the segment pcn is the diameter. Therefore given a fixed A, the normal section has the smalles curvature and curvatures of all remaining plane sections increases in the way described in the Meusnier theorem. JS: missing picture An example if the sphere where these sections are circles with the radius which decreases in this way. 58 8.18 8.18. Finally we consider a a class of surfaces which is interested both geometrically and from a point if view of applications. Definition. The surface S is called minimal if its mean curvature H is zero in all points. A nontrivial example of a minimal surface is the helicoid which we shall study in sections 10.7 and 10.9. The adjective “minimal” is based on the variational calculus. One of important variational problems is the problem to “span” a surface with a minimal are to the given curve in E3. Under rather general conditions, solutions of this problem are surfaces with minimal mean curvature. 59 9 Envelope of a family of surfaces 60 10 Ruled surfaces 10.1. One-parameter system of lines in E3 is a mapping which each t ∈ I maps to a line p(t) where I is an open interval. The line p(t) is called generating line of the surface. This line can be described using a point g(t) ∈ p(t) and a nonzero vector h(t) in the direction of p(t). An arbitrary point of the line p(t) then has the form e10.1e10.1 (1) f(t, v) = g(t) + vh(t) , v ∈ R . Similarly as in agreement 1.15 or 4.7, we shall further assume g(t) and h(t) are functions of the class Cr where r is big enough for our consideration. Then f : I × R → E3 is map of the class Cr. Thus, in a sense, this is a two parametric movement. The condition from the definition of surfaces requires, beside injectivity of f, also vectors ∂f ∂t := ft = g + vh and ∂f ∂v := fv = h to be lienarly independent at every point. We shall illustrate this on examples. 10.2 10.2. Let the point g(t) = a be fixed. Then we have generalized cone e10.2e10.2 (2) f(t, v) = a + vh(t) . Therefore ft = vh , fv = h, ft×fv = v(h ×h). Tha value v = 0 corersponds JS: missing picture to the vertex of the cone which is obviously a singular point. Given v = 0, we need h ×h = o for all t ∈ I. If this is satisfied then, assuming injectivity of f, we get a surface. If h × h = 0 everuwhere, we have e10.3e10.3 (3) ∂h ∂t = k(t) h(t) where k(t) is a reak function. If we consider a real function z(t) instead of z(t) then our differential equation, by separation of variables, has the general solution z = l(t)c where l(t) = eSk(t) dt. This holds for each component of our vector valued function h(t) hence h(t) = l(t)b where b is a constant vector. That is, this is the case of a two parametric movement along a line a not a surface. 10.3 10.3. Let h(t) = a is a fixed nonzero vector. Then we get generalized cylinder JS: missing picture e10.4e10.4 (4) f(t, v) = g(t) + va , t ∈ I, v ∈ R . Here ft = g , fv = a. If g (t) × a = o for all t and f is injective, it is a surface. If the tangent vector g (t) is colinear with the vector a at some point, this point is singular. 61 10.4 10.4. Consider a curve C ≡ g(t) given parametrically and consider the tangent line at every point g(t). This one parametric family of curves is called tangent developable of the curve C. Its parametric expression has the form e10.5e10.5 (5) f(t, v) = g(t) + v g (t) . We have ft = g (t) + v g (t), fv = g (t) hence e10.6e10.6 (6) ft × fv = −v g (t) × g (t)) . Further we assume that C does not hav inflection points Then the vector (6) is zero if and only if v = 0 which is the original point on the curve. Consider the normal plane ν(t, v) at the point g(t0) of the curve C. Using scalar product, this is given by e10.7e10.7 (7) g (t0), w − g(t0) = 0 , w(x, y, z) ∈ E3 . The tangent line at g(t) is given parametrically as g(t) + v g (t) . Denote by v(t) the parameter of its intersection with ν(t0). That is, e10.8e10.8 (8) g (t0), g(t) + v(t) g (t) − g(t0) = 0 . Such intersection is a movement in the normal plane ν(t0, 0) given by the parametrization e10.9e10.9 (9) h(t) = g(t) + v(t)g (t) , v(t0) = 0 . We shall sho that h (t0) = o. We have h (t0) = g (t0) + v (t0) g (t0) + v(t0) g (t0) . Since v(t0) = 0, it is enough to prove that v (t0) = −1. By differentation of (8) we get g (t0), g (t) + v (t) g (t) + v(t) g (t) = 0 . Putting t = t0, we obtain g (t0), g (t0) + v (t0) g (t0), g (t0) = 0 . 62 Since g (t0), g (t0) = 0, we have v (t0) = −1. Considering the movement h(t), the condition h (t0) = o means that h(t0) is a singular point. Generally, it is an edge, cf. 1.9. It is easier to think about the tangent developable of the screw line whose projection in the JS: missing picture direction of the axes of the screw line is on the picture. This geometrically shows that the tangent developable is not a curve in the sense od the definition 4.6 in a neighbourhood of the generating curve. 10.5 10.5. If one wants to obey the definition 4.6 also in the case of oneparameter systems of lines, the following approach should be used: Definition. A surface S ⊂ E3 is called ruled surface if S is a part of a one-parameter system of lines. We shall talk about a generating line of the surface S in such case although it can be only a part of a line. 10.6. We shall further discuss two examples. First consider 3 skew lines q1, q2, q3 (i.e. 3 lines in general position). We consider the transversal of skew lines q1, q2 through every point a ∈ q3 (this is the intersection of planes given by the point p and either the line q1 or q2). It is shown in theory of the projective geometry that this construction gives rise to a regular ruled quadric. Starting with 3 lines of the system of lines we have just constructed and repeating this construction, we obtain the second JS: missing picture one-parameter system of lines on the same regular ruled quadric. 10.7 10.7. Given a ruled surface which is neither a regular ruled quadric nor a part of a plane, we have shown there can be, beside generating lines, at most two more lines. This is related to the followng general construction. Chose two skew lines q1, q2 and a curve C. We consider the transversal of JS: missing picture skew lines q1, q2 through every point a ∈ C. This yields a one-parameter system of lines An important example (for technical applications) is the case when one of skew lines is an improper line of some plane . Thus we have the plane JS: missing picture , a line q and a curve C. We consider lines intersecting q and C which are parallel with . This yields a one-parameter system of lines which is called conoid. If the line q is perpendicular to the plane , we obtain right conoid. Example. IF C is the screw line, q its axis and a plane perpendicular to q, the we obtain so called right screw conoid or helikoid. We mentioned JS: missing picture 63 this surface in 8.18. Consider q to be z-axis and C to lie on the the cylinder with he unit radius around the z-axis. Parametric description of C then is (cos t, sin t, bt), b = 0, t ∈ R. The helicoid has the form e10.10e10.10 (10) f(t, v) = (v cos t, v sin t, bt) , b = 0, v ∈ R . It is easy to see that kinematically this surface is built by screwing a generating line perpendicular to and intersecting the axis q, in the direction of this axis. de10.8 10.8 Definition. A ruled surface is called developable if tangent planes at all ponits of a fixed generating line are the same. We say the tangent plane is tangent along generating lines. Geometrically it is clear (and the related computation is easy) that this is a property of generalized cylinders and cones. We shall show that also in also in the case of tangent developable of the curve C, the tangent line of the surface is the same at all point of a generating line. Following 10.4, the tangent plane of the surface g(t) + vg (t) at the point g(t0) + v0g (t0), v0 = 0 is determined by this point and vectors g (t0) and g (t0). For a fixed t0 and each v0 = 0, this plane coincides with the osculating plane of the curve C at the point g(t0) + vg(t0). On the other hand, the tangent plane along a generating line of a regular quadric is not fixed. Here the tangent plane is determined by the given generating line and by a line of the other system through a given point. However, the other system is formed by trasnversal lines which cannot lie in a fixed plane. Thus considering the helicoid from 10.7, the tangent plane changes along a generating line. It follows from (10) that e10.11e10.11 (11) ft = (−v sin t, v cos t, b) , fv = (cos t, sin t, 0) thus the unit normal vector is ft × fv ft × fv = 1 √ v2 + b2 (−b sin t, b cos t, −v) which changes depending on v (for a fixed t = t0). 10.9 10.9. Further we shall need to express coefficients of the second fundamental form using the exteriorproduct We know the exterior product [a, b, c] of three vectors in the euclidean oriented three-dimensional space is formed id equal to the scalar product of the vector product of first vectors with the third vector, i.e. [a, b, c] = (a × b, c) . 64 Using this to formulae 7.(1) and 7.(5), we obtain e10.12e10.12 (12) h11 = 1 f1 × f2 [f1, f2, f11] , h12 = 1 f1 × f2 [f1, f2, f12] , h22 = 1 f1 × f2 [f1, f2, f22] . Example. We shall show the helicoid from (10) is a minimal surface in the sense of 8.18, i.e. that H = 0. Differentiating of (11) we obtain ftt = (−v cos t, −v sin t, 0), ftv = (− sin t, cos t, 0) and fvv = o. Thus h11 = 0 and h22 = 0. Since also g12 = 0, the formula 8.(15) yields H = 0. 10.10 10.10. Since every line on a surface is asymptotic curve, all points on ruled surfaces are hyperbolic, parabolic or planar. Theorem. A developable ruled surfaces S has zero Gaussian curvature. Proof. Given a parametrization f(t, v) = g(t) + v h(t) of the surface plochy S, we have f1 = g (t) + v h (t), f2 = h(t). For a fixed t, the vector h(t) lies in the associated vector space of the tangent plane for all v if for the vectors h(t) and g (t)+v1h (t), g (t)+v2h (t) are complanar for all v1 = v2. By a lienar combination of these vectors we obtain g (t) and h (t) hence the condition for a fixed tangent plane is e10.13e10.13 (13) g (t), h(t), h (t) = 0 . A further computation yiedls f11 = g (t) + v h (t) and f12 = h (t), f22 = o. Following (12) we obtain h22 = 0, h12 = 1 f1 × f2 g (t) + v h (t), h(t), h (t) and also h12 = 0 according to (13). It follows from 8.(15) that K = 0 independently on h11. 10.11 10.11. In the opposite direction we have the following statement. Theorem. If every point of the surface S is parabolic then S is locally a developable ruled surface. 65 Proof. Given a parabolic surface S, we chose local paramatrization such that the net of asymptotic curves is given by u1 = konst.. Then 0 = h12 = (n, f12) and 0 = h22 = (n, f22). We know that (n, f1) = 0, (n, f2) = 0, (n, n) = 1. Differentiating this with respect to u2 yields – similarly as in 7.(8) – that (n2, f1) = 0 , (n2, f2) = 0 , (n2, n) = 0 . Thus n2 = o, i.e.. the normal vector along the curve u1 = konst. is fixed. Followign Podle 7.(8), it follows from h12 = 0 that (n1, f2)=0. Differenting this with respect to u2 and using u12 = o, we obtain (n1, f22 = 0). Hence vectors f2 and f22 are perpendicular to vectors n and n1 which are linearly independent. Indeed, the vector n1 is perpendicular to n and is nonzero. Indeed, the first of equations in 7.(8) means (n1, f1) + (n, f11) = 0. Thus n1 = o would mean h11 = 0 which, together with h12 = 0 and h22 = 0, would identify a planar point which we exclude. Sicne vectors f2 and f22 are perpendicular to two linearly independent vectors, they are colinear. Hence every point of the curve u1 = konst. is inflection, i.e. this curve is a line. The tangent line along this generating line is constant thus S is locally a developable ruled surface. de10.12 10.12 Definition. Generating line g(t0) + vh(t0) of a ruled surface (1) is called cylindric if vectors h (t0) and h(t0) are colinear. Theorem. A ruled surface with all generating lines cylindric is a generalized cylinder. Proof. We have shown in 10.2 that it follows from dh dt = k(t) h(t) that h(t) = l(t)b where b is a constant vector. Thus we have f(t, v) = g(t) + v l(t)b which is a cylinder with a different parametrization of generating lines. 10.13 10.13. Similarly, we have a direct geometrical characterization of a cylindrical line. Given a parametrization p(t) ≡ g(t) + v h(t), we can assume that the vector h(t) is unit. Then h(t) is a motion along a unit sphere which is called spherical image of a ruled surface. Differentiating (h, h) = 1 we obtain (h, h ) = 0 thus the vector h (t) is perpendicular to h(t) for eveery t. Assuming further that h (t0) is colinear with h(t0), we obtain h (t0) = o. Therefore 66 Theorem. Generating line p(t0) of a ruled surface S is cylindric if and only if t0 is a singular point of the sphercial image of the surface S. 10.14 10.14. The meaning of the word “generally” in the following statement is explained during the proof. Theorem. A ruled surface without cylindrical lines is generally either a tangent developable or a cone. Proof. 10.15 10.15. Our results from 10.12 and 10.14 are sometimes summarized as surface with zero Gaussian curvature is generally either a tangent developable or a generalized cylinder or a generalized cone. 67 11 Isometric mappings 11.1 11.1. Let D, ¯D ⊂ R2 be open sets. A map ϕ: D → ¯D is given by a pair of functions ϕ1, ϕ2 : D → R which are called components of the map ϕ. Denoting u1, u2 coordinates in D and v1, v2 coordinates in ¯D, we have v1 = ϕ1(u1, u2), v2 = ϕ2(u1, u2). Definition. We say ϕ is mapping of the class Cr if ϕ1 and ϕ2 are functions of the class Cr. Henceforth we assume ϕ is a mapping of the class Cr, r ≥ 1. de11.2 11.2 Definition. The determinant J(ϕ) = ∂ϕ1 ∂u1 , ∂ϕ1 ∂u2 ∂ϕ2 ∂u1 , ∂ϕ2 ∂u2 is called Jacobian of the mapping ϕ. 11.3 11.3. We shall study mapppings g: S → ¯S between two simple surfaces of the class Cr. Assume S and ¯S are given by parametrizations f(u1, u2), JS: missing picture (u1, u2) ∈ D and ¯f(v1, v2), (v1, v2) ∈ ¯D. The map g determines a unique map ψ: D → ¯D such that g◦f = ¯f ◦ψ. Thus ψ expresses g in the parameter space. In the opposite direction, ψ deteremines the mapping g. Definition. We say that g: S → ¯S is mapping of the class Cr if the corresponding mapping ψ: D → ¯D is of the class Cr. 11.4 11.4. A choice of parametrizations of surfaces S and ¯S is used in the previous definition. We shall show that the class of differentiability of the mapping g is independent on the choice of parametrizations f and ¯f. First we discuss a change of the parametrization f(u1, u2) of the surface S, (u1, u2) ∈ D. Consider a bijective mapping ϕ: ¯D → D of the class Cr where ϕ = ϕ1(v1, v2), ϕ2(v1, v2) , (v1, v2) ∈ ¯D. Then ¯f = f ◦ ϕ is also of the class Cr. In the case of parametrziation of a surface we have further the condition f1 × f2 = o. Using the notation ¯f1 = ∂1(f ◦ ϕ), ¯f2 = ∂2(f ◦ ϕ) and using the chain rule, we obtain e11.1e11.1 (1) ¯f1 = f1 ∂ϕ1 ∂v1 + f2 ∂ϕ2 ∂v1 , ¯f2 = f1 ∂ϕ1 ∂v2 + f2 ∂ϕ2 ∂v2 . Thus ¯f1 × ¯f2 = ∂ϕ1 ∂v1 ∂ϕ2 ∂v2 − ∂ϕ2 ∂v1 ∂ϕ1 ∂v2 f1 × f2 = J(ϕ) f1 × f2 . 68 Definition. A bijective map ϕ: ¯D → D of the class Cr is called reparametrization if J(ϕ) = 0for all (v1, v2) ∈ ¯D. 11.5 11.5. Now it is clear that the definition 11.3 does not depend on the choice of parametrization. Consider a reparametrization ϕ: D1 → D on S and a reparametrizaci ¯ϕ: ¯D1 → ¯D on ¯S. Then ¯ϕ−1 : ¯D → ¯D1 is also a map of the class Cr (this follows immediately from the generalized implicit function theorem, cf. e.g. the textbook [5]). Thus we have ¯ϕ−1 ◦ ψ ◦ ϕ instead of the mapping ψ in the definition 11.3 which is also of the class Cr. 11.6 11.6. Further we shall consider motion on surfaces. First consider the map ψ: D → ¯D, v1 = ψ1(u1, u2), v2 = ψ2(u1, u2). Then D itself is a surface (as a part of plane) hence we have the tangent plane TuD for every u ∈ D which coincides with R2. Its elements are tangent vectors to motions h(t), h: I → D at the point h(0) = u where we assume 0 ∈ I. Coordinates of the tangent vector are dh1(0) dt , dh2(0) dt . Consider the motion ψ ◦h: I → ¯D. Coordinates of its tangent vector at t = 0, which one findss applying the chain rule to ψ1 h1(t)h2(t) and ψ2 h1(t), h2(t) , are e11.2e11.2 (2) ∂ψ1 ∂u1 dh1(0) dt + ∂ψ1 ∂u2 dh2(0) dt , ∂ψ2 ∂u1 dh1(0) dt + ∂ψ2 ∂u2 dh2(0) dt . That is, the tangent vector d(ψ◦h)(0) dt is determined by the vector dh(0) dt . Considering the linearity of (2), we have Theorem. The rule dh(0) dt → d(ψ◦h)(0) dt deteremines a linear mapping Tuψ: TuD → Tψ(u) ¯D for each u ∈ D. Definition. This mapping is called tangent mappping ψ at the point u. Denoting by (du1, du2) coordinates at TuD by (dv1, dv2) coordinates in Tψ(u) ¯D then (2) has the form e11.3e11.3 (3) dv1 = ∂ψ1 ∂u1 du1 + ∂ψ1 ∂u2 du2 , dv2 = ∂ψ2 ∂u1 du1 + ∂ψ2 ∂u2 du2 . Thus these are differentials of functions ψ1 and ψ2. 11.7 11.7. Consider the original map g: S → ¯S and put p = f(u) ∈ S. Consider the tangent space TpS and a vector vektor A ∈ TpS which is tangent to the motion γ(t) on S, A = dγ(0) dt . Then g ◦ γ is a motion on ¯S and the tangent vector d(g◦γ)(0) dt to this motion depends only on A. Indeed, it is given (2) in terms of the parametrization. Thus we have shown 69 Theorem. The rule dγ(0) dt → d(g◦γ)(0) dt determines a linear map Tpg: TpS → Tg(p) ¯S. Definition. The map Tpg is called tangent map to g at the point p. de11.8 11.8 Definition. We say the mapping g: S → ¯S is isometric if each tangent map Tpg: TpS → Tp ¯S, p ∈ S preserves the scalar product. That is, (A, B) = Tpg(A), Tpg(B) for each A, B ∈ TpS. If g is bijective, it is isometry of S and ¯S. 11.9 11.9. The bijiective map g: S → ¯S can be realised in sucha way we cos- JS: missing picture nider a common parameter space D and corresponding points f(u1, u2) ∈ S and ¯f(u1, u2) ∈ ¯S. In this case we say the map g is given by equality of parameters. Theorem. The bijection g: S → ¯S given by equality of parameters is isometry if and only if first fundamental forms Φ1 and ¯Φ1 of surfaces S and ¯S, respectively are coincide. Proof. Vectors f1, f2 form a basis in TpS, vectors ¯f1, ¯f2 form a basis in Tg(p) ¯S and the linear map Tpg is given by du1 = du1, du2 = du2. Following 6.1, scalar products of tangent vectors to S and ¯S are given by the first fundamental form. The condition Φ1 = ¯Φ1 explicitly means g11 = ¯g11, g12 = ¯g12, g22 = ¯g22 where bared quantities are computed on the surface ¯S depending on the same parameters (u1, u2). 11.10 11.10. The following statement justifies the terminology of isometry. Theorem. The bijection g: S → ¯S is isometry if and only if it preserves length of curves. Proof. The bijection g can be given by equality of parameters. Since the length of the curve u1(t), u2(t) , t ∈ [a, b] is e11.4e11.4 (4) s = b a g11 du1 dt 2 + 2g12 du1 dt du2 dt + g22 du2 dt 2 dt , it follows from the theorem 11.9 that length of curves corresponding to ismoetries are the same. In the opposite direction, if both curves with the same parametric expression have the same length, also their tangent vectors have the same length according to (4). Thus the linear map Tpg for each p ∈ S preserves length of vectors. Now it follows from linear algebra theory that such mappresrves also the scalar product. 70 11.11 11.11. Geometrically it is obvious that the cylinder f(u) = (r cos u1, r sin u1, u2), u1 ∈ (0, 2π), u2 ∈ R is isometric with a plane strip with the same coordinates. Physically, this isometry is given by unfolding the cylinder into the plane. We verify this also by computation using the theorem 11.9. We have f1 = (−r sin u1, r cos u1, 0), f2 = (0, 0, 1) hence g11 = r2, g12 = 0, g22 = 1. The plane z = 0 can be parametrized as ¯f(u) = (ru1, u2, 0). Thus ¯f1 = (r, 0, 0), ¯f2 = (0, 1, 0) and ¯g11 = r2, ¯g12 = 0, ¯g22 = 1 which is the same as in the case of cylinder. de11.12 11.12 Definition. By inner geometry of the surface S we mean properties preserved by isometries. It follows from the theorem 11.9 that inner geometry of the surface is formed by properties which can be derived from the first fundamental form. We say that properties of S which essentially depend on the second fundamental form, belong to outer geometry of the surface. 11.13 11.13. The notion of inner geometry of the surface originates in Gauss’ work. He derived the following its most important statement. It is a deep result which will be proved in the next section in 12.14. Gauss called this Teorema egregium in latin (the transaltion: remarkable theorem). Let KS be the Gaussian curvature of the surface S K¯S the Gaussian curvature of the surface ¯S. Teorema egregium (Gauss). If g: S → ¯S is an isometry then KS(p) = K¯S g(p) for all p ∈ S. One should emphasize that both principal curvatures do not belong to inner geometry of the surface (the second fundamental form is essential for their computatin) whereas their product does. 11.14 11.14 Example. An open set in the plane cannot be isometric to an open set on the sphere. Indeed, The Gaussian curvature of the plane is zero whereas the Gaussian curvature of the sphere of the radius r is 1 r2 . 11.15 11.15. Recall by neighbourhood on the surface S of the point p ∈ S we mean intersection of a neighbourhood of p in E3 with the surface S. Definition. The surface S is called developable if each point p ∈ S has a neighbourhood isometric to an open set in the plane. This isometry is understood as “unfoldability” of the corresponding part of the surface into the plane, see the example 11.11 which justifies this terminology. On the other hand, we introduced the notion of a developable 71 ruled surface at 10.8. The same terminology is based on the fact that both notions coincide as we shall show. If necessary to distinguish both definitions, the latter one will be referred as developability in a metric sense. 11.16 11.16. Consider a generalized cylinder. Assume the z-axis is parallel with generating lines and the curve g is the section of the cylinder by the plane z = 0 parametrized by the arc-length. The the cylinder is given by parametrization f(s, v) = g1(s), g2(s), v . Thus f1 = (g1, g2, 0), f2 = (0, 0, 1), g11 = 1, g12 = 0, g22 = 1. Using the parametrization ¯f(s, v) = (s, v, 0) of the plane, we obtain the same coefficients ¯g11 = 1, ¯g12 = 0, ¯g22 = 1. Hence equality of parameters yields an isometry of both surfaces. 11.17 11.17. Consider a generalized cone with the vertex at the origin. Thus f(t, v) = g(t) v. Here we can assume the curve g is parametrized by arclength s and g(s) = 1 for all s. Then f1 = g (s) v, f2 = g(s), i.e. g11 = v2, g22 = 1 and g12 = 0 since vectors g(s) and g (s) are perpendicular. If we choose g as the unit circle k(s) in the plane we can parametrized the plane JS: missing picture locally in the form ¯f(s, v) = k(s) v. Hence ¯g11 = v2, ¯g12 = 0, ¯g22 = 1 which shows a local isometry of the plane and the cone. 11.18 11.18. Consider the tangent developable g(t) + vg (t) of the curve C. We assume C is parametrized by arc-length. Using Frenet formuale, we can parametrize the surface as f(s, v) = g(s) + ve1(s) . Thus f1 = e1(s) + vκ(s) e2(s), f2 = e1(s), i.e. g11 = 1 + v2 κ2(s), g12 = 1, g22 = 1. Further consider a curve ¯g(s) in the plane which has locally the same curvature as the spacial curve C and parametrize the plane locally as ¯f(s, v) = ¯g(s) + v ¯g (s). This can be written also in the form ¯f(s, v) = ¯g(s) + v ¯e1(s) . We have ¯f1 = ¯e1(s) + v κ(s) ¯e2(s), ¯f2 = ¯e1(s). Also in this case we have ¯g11 = 1 + v2κ2(s), ¯g12 = 1, ¯g22 = 1. Thus we obtained a local isometry of teh tangent developable with the plane. 11.19 11.19. We have shown in 10.11 that a surface with zero Gaussian curvature without planar points is locally a developable ruled surface. We know 72 from 10.15 that a developable ruled surface is generally either tangent developable orgeneralized cylinder or generalized conev. We have shown these surfaces are developable in the metric sense. Using an alternative approach (which we shall not discuss here) can be shown the following statement (cf. the theorem 14.6). Theorem. A surface S is locally isometric to the plane if and only if it has zero Gaussian curvature. 73 12 Parallel transport of vectors on a surface 12transport 12.1 12.1. Consider a surface S ⊂ E3 and a motion γ : I → S. We denote by V the associated vector space of E3. Definition. The mapping A: I → V is callede tangent vector field on S aling the motion γ if A(t) ∈ Tγ(t)S for all t ∈ I. Zero tangent vectors along γ form the field denoted by 0γ. If A is a tangent vector field along γ and g: I → R is a function then g(t)A(t) is also a tangent vector field along γ. If A1 and A2 are two tangent vector fields along γ then also A1(t) + A2(t) is a tangent vector field along γ. 12.2 12.2. Recall NpS is the normal vector space of the surface S at the point p. The following definition is essential for the differential geometry of surfaces. Definition. We say that tangent vectior field A along the motion γ is formed by vectors parallel on S if dA dt ∈ Nγ(t)S for all t ∈ I. 12.3 12.3. The vector dA dt decomposes into a direction in the tangent plane Tγ(t)S and the normal line Nγ(t)S at each point of the motion γ(t). TTangent components form again a tangent vector field on S along γ. Definition. Tangent vectir field A dt on S along γ formed by tangent components of vectors dA dt is called textbfcovariant derivative of the tangent vector field A on S along the motion γ. Thus the field A is formed by vectors parallel on S if and only if A dt is the zero field along γ. 12.4 12.4. We shall find a parametric expression for A dt . Since vectors f1, f2, n are linearly independent at each point of the surface, we can decompose second partial derivatives as e12.1e12.1 (1) f11 = Γ1 11(u1, u2) f1 + Γ2 11(u1, u2) f2 + h11(u1, u2) n , f12 = Γ1 12(u1, u2) f1 + Γ2 12(u1, u2) f2 + h12(u1, u2) n , f22 = Γ1 22(u1, u2) f1 + Γ2 22(u1, u2) f2 + h22(u1, u2) n . Taking scalar product of each equation with the unit vector n perpendicular to f1 and f2 shows that coefficients at n are indeed coefficients of the second fundamental form as the notation indicates. 74 Definition. The function Γi jk, i, j, k = 1, 2, Γi 21 = Γi 12 are called Christoffel symbols of the surface S corresponding to the parametrization f(u1, u2). 12.5 12.5. Let the motion γ(t) is given by the parametrization u1(t), u2(t) and A(t) = U1(t), U2(t) . Then e12.2e12.2 (2) A(t) = U1(t) f1 u1(t), u2(t) + U2(t) f2 u1(t), u2(t) . From this one directly computes using (1) that dA dt = dU1 dt f1 + U1 f11 du1 dt + f12 du2 dt + dU2 dt f2 + U2 f12 du1 dt + f22 du2 dt = dU1 dt + Γ1 11U1 du1 dt + Γ1 12 U1 du2 dt + U2 du1 dt + Γ1 22U2 du2 dt f1 + dU2 dt + Γ2 11U1 du1 dt + Γ2 12 U1 du2 dt + U2 du1 dt + Γ2 22U2 du2 dt f2 + (. . . )n . where an expression at n is not important for us. Denoting U1 dt , U2 dt coordinates of the vector field A dt along γ, we have e12.3e12.3 (3) U1 dt = dU1 dt + 2 i,j=1 Γ1 ij u(t) Ui duj dt , U2 dt = dU2 dt + 2 i,j=1 Γ2 ij u(t) Ui duj dt . 12.6 12.6. As the first property of formulae (3) we shal derive th following: Theorem. Tangent vector fields A, B on S along the motion γ and the function g: I → R satisfy e12.5e12.5 (4) (A + B) dt = A dt + B dt , (gA) dt = dg dt A + g A dt . Proof. This follows directly from (3). 12.7 12.7. The following statement shows that the parallel transport along a given motion has similar properties as the parallel transport of vectors in the plane. Theorem. For each motion γ : I → S, each t0 ∈ I and each vector A0 ∈ Tγ(t0)S exists unique vectir field along γ satisfying A(t0) = A0 and formed by vectors parallel on S. 75 Proof. Given the motion γ, the condition that right hand sides of expressions of (3) foms a system of two ordinary differential equations. The value A0 is the initial condition for this system. This determines the solution uniquelly. de12.8 12.8 Definition. We say the tangent vector field A from the theorem 12.7 provides parallel transport of the vector A along the motion γ on S. Assume the motion γ(t) is the parametrized simple curve C on S. A reparametrization t = ϕ(τ) of C yields another motion γ ◦ ϕ. Inserting this reparametrization into (3), derivation with respect to t are mutliplied by dϕ dτ . Since expressions (3) are linear in dUi dt a dui dt , i = 1, 2, differential equations Ui dt = 0, i = 1, 2 for parallel transport are essentially the same as (Ui◦ϕ) dτ = 0. Geometrically this means for different parametrizations of the curve C on S we get the same parallel transport of vectors. Thus we speak not only about paralle transport of vectors along a motion on S but also about parallel transport of vectors along a curve on S. 12.9 12.9 Theorem. If tangent vecros fields A and B along the motion γ provide parallel transport of vectors A0 ∈ Tγ(t0)S and B0 ∈ Tγ(t0)S, respectively, then the field k1A +k2B provides the parallel transport of the vector k1A0 + k2B0, k1, k2 ∈ R. Proof. The theorem 12.6 for constant g = k yields (kA) dt = k A dt . Thus (k1A+k2B) dt = k1 A dt + k2 B dt . Since the right hand side is zero, the same holds for the left hand side. Geometrically this means the parallel transport preserves lienar combinations of vectors. 12.10 12.10 Example. Considering a plane as a surface in E3 then for the tangent vector field A(t) = U1(t), U2(t) along an arbitrary motion γ(t) in has zero normal component of the vector dA dt . Thus A(t) is parallel transport along γ if and only if dA dt = 0, i.e. U1(t) a U2(t) are constants. This is the classical parallel transport in the plane. This transport does not depend on the motion. Now we shall show that, however, parallel transport on the sphere S does depend on the motion. Conside one eighth of the sphere according to the picture. The great circle in the plane z = 0 has the paramet- JS: missing picture ric expression f(t) = (r cos t, r sin t). Its tangent vector is v(t) = df dt = 76 (−r sin t, r cos t). Thus dv dt = (−r cos t, −r sin t). The normal vector of the sphere at the point f(t) is (cos t, sin t) hence dv dt ∈ Nf(t)S. Therefore tangent vectors to the great circle provide transport parallely along this circle. Denote by a the point with the parameter t = 0 and b the point with the parameter t = π 2 . Now let us transport this tangent vector v = v(0) along the great circle in the plane perpendicular to v at the point c with the parameter π 2 . The constant vector v is tangent along this curve, i.e. dv dt = o. Now let us continue with the parallel transport along the small arc of the great circle from the point c to b. Here we again transport the tangent vector of the circle along this circle hence the result is again tangent to this circle. Thus the resulting transport of the vector v from the point a to b along two different motions I and II give rise to two different vectors which are actualy perpendicular. Although the second motion is not differentiable at one point, the “vertex” at the point c is not the reason of different parallel transports of the vector v along motions I and II. 12.11 12.11. Decompositions (12.1) can be used also to compute Christoffel symbols. This is particularly simple in the case when the parametric net is orthogonal, i.e. g12 = (f1, f2) = 0. Let us discuss the sphere f(u1, u2) = r(cos u1 cos u2, sin u1 cos u2, sin u2) from 6.4. We found f1 = r(− sin u1 cos u2, cos u1 cos u2, 0) , f2 = r(− cos u1 sin u2, − sin u1 sin u2, cos u2) ,(5) g11 = (f1, f1) = r2 cos2 u2 , g12 = (f1, f2) = 0 , g22 = (f2, f2) = r2 . Further differentiation yields e12.6e12.6 (6) f11 = r(− cos u1 cos u2, − sin u1 cos u2, 0) , f12 = r(sin u1 sin u2, − cos u1 sin u2, 0) , f22 = r(− cos u1 cos u2, − sin u1 cos u2, − sin u2) . We consider scalar product of equations (1) with vectors f1 and f2 where we need to compute scalar products on the left hand side. We obtain 0 = Γ1 11r2 cos2 u2, r2 sin u2 cos u2 = Γ2 11r2 , −r2 sin u0 cos u0 = Γ1 12r2 cos2 u2 , 0 = Γ2 12r2 , 0 = Γ1 22r2 cos u2 , 0 = Γ2 22r2 . Thus Γ1 11 = 0 , Γ2 11 = sin u2 cos u2 , Γ1 12 = − tan u2 , Γ2 12 = 0 , Γ1 22 = 0 , Γ2 22 = 0 . 77 12.12 12.12. The following theorem is an important tehoretical result. Equations (12.7) show that Christoffel symbols can be expressed using coefficirnts of the first fundamental form. THUS THE PARALLEL TRANSPORT OF VECTOS ON THE SURFACE BELONGS TO THE INNER GEOMETRY OF THE SURFACE. Since g11g22 −g2 12 > 0, the square matrix (2×2)-matrix (gij) is reguklar. Denote by (gkl) its inverse matrix. Theorem. We have e12.7e12.7 (7) Γk ij = 1 2 2 l=1 gkl ∂gjl ∂ui + ∂gli ∂uj − ∂gij ∂ul . Proof. Differentiating (f1, fj) = gij with respect to ul, we obtain e12.8e12.8 (8) ∂gij ∂ul = (fil, fj) + (fi, fjl) . It follows from (1) that fij = 2 m=1 Γm ij fm + hijn. By substitution we obtain (fil, fj) = 2 m=1 Γm il gmj . Thus (8) can be written as e12.9e12.9 (9) ∂gij ∂ul = 2 m=1 Γm il gmj + Γm jl gmi . We obtain two more equations by a suitable change of indices, Dal? dv? rovnice z?sk me z m?nou index? ∂gil ∂uj = 2 m=1 Γm ij gml + Γm lj gmi ,e12.10e12.10 (10) ∂glj ∂ui = 2 m=1 Γm ji gml + Γm li gmj .e12.11e12.11 (11) Summing (10) + (11) − (9) and using symmetris of gij and Γk ij at lower indices, we obtain e12.12e12.12 (12) ∂gil ∂uj + ∂glj ∂ui − ∂gij ∂ul = 2 2 m=1 Γm ij gml . 78 By division by 2 and for fixed i, j, we consider (Γm ij ) as rank two row vector. Then the matrix form of the right hand side of (12) is (Γm ij )(gml). Now we compute the unkown Γm ij by multiplication of the inverse matrix (gkl). This yields (7). 12.13 12.13. Of course, (7) can be used also as a formula to compute Christoffel symbols. A simple example is the generalized cylinder from 11.16. where we found g11 = 1, g12 = 0, g22 = 1. All partial derivatives in (7) are thus zero since these are derivatives of a constant. Hence all Christoffel symbols of the generalized cylinder are zero. 12.14 12.14. Now we shall prove the Gauss’ theorem egregium. We shall start with equations (1) which we shall wwrite in the form e12.13e12.13 (13) fij = 2 l=1 Γl ijfl + hijn . We shall use the notation fijk = ∂3f ∂ui∂uj∂uk , Γl ij,k = ∂Γl ij ∂uk , ni = ∂n ∂ui , hij,k = ∂hij ∂uk . Equations 7.(8) can be written in the form e12.14e12.14 (14) (ni, fj) = −hij . Differentiating (13) with respect to uk, we obtain e12.15e12.15 (15) fijk = 2 m=1 Γm ij,kfm + 2 n=1 Γn ijfnk + hij,kn + hijnk . Using (13), the second term on the right hand side has the form e12.16e12.16 (16) 2 n=1 Γn ijfnk = 2 m,n=1 Γn ij(Γm nkfm + hnkn) . Taking the scalar product of (15) with the vector fj and using (14), we obtain e12.17e12.17 (17) (fijk, fl) = 2 m=1 Γm ij,kgml + 2 m,n=1 Γn ijΓm nkgml − hijhkl . It follows from symmetry of third partial derivatives fikj = fijk that (17) holds also after exchange of indices j and k. Subtracting these two equations, we obtain e12.18e12.18 (18) hijhkl − hikhjl = 2 m=1 gml Γm ij,k − Γm ik,j + 2 n=1 (Γn ijΓm nk − Γn ikΓm nj) . 79 Putting i = 1, j = 1, k = 2, l = 2 we obtain Theorem (Gauss’ equations). It holds e12.19e12.19 (19) h11h22 − h2 12 = 2 m=1 gm2 Γm 11,2 − Γm 12,1 + 2 n=1 (Γn 11Γm n2 − Γn 12Γm n1) . The right hand side depends only on coefficients gij of the first fundamental form and its partial derivatives of thr first and second order according to the theorem 12.12. We have found in 8.9 that K = (h11h22 − h2 12)/(g11g22 − g2 12). Thus the Gaussian curvature K belongs to the inner geometry of the surface as the theorema egregium states. We also remark that putting different indices i, j, k, l into (18), we obtain again the equation (19) or identity. 12.15. Information If we analogously decompose ni, i = 1, 2 into the12.15 frame f1, f2, n and use the relation ∂ni ∂uj = ∂nj ∂ui = ∂2n ∂ui∂uj , we obtain so called Codazzi equations (two of them are essentially) e12.20e12.20 (20) hij,k − hik,j + 2 l=1 (Γl ijhlk − Γl ikhlj) = 0 . Using elementary techniques for systems of partial differential equations (e.g. the theorem 7.8 in the textbook [5]), one can prove following two statements about existence and uniqueness of solutions, which are sometimes called The basic theorem of theory of surfaces. I. If S and ¯S are two simple surfaces with parametrizations f : D → E3 and ¯f : D → E3, respectively on the same parameter space D, which have the same first and second fundamental form, then there exists an Euclidean motion ϕ: E3 → E3 such that ϕ ◦ f = ¯f. Summarizing: surfaces, which have the same first and second fundamentzal form, are congruent. II. Consider two quadratic forms Φ1 and Φ2 on D ⊂ R2 where Φ1 is positive definite at all points. If Φ1 and Φ2 satisfy Gauss and Codazzi equations, then there locally exists a surface with a parametrization f : D → E3 such that Φ1 and Φ2 are its first and second fundamental form, respectively. 80 13 Geodetic curves 13 13.1 13.1. Given a surface S and a motion γ : I → S, we can consider the field of tangent vectors of γ which we denote by ˙γ. Definition. The motion γ : U → S is called geodetic curve if the field ˙γ of its tanegnt vectors transports parallely along γ. Consider a plane as a surface, this definition means that the vector ˙γ = a is constant, i.e. γ is the motion p + ta along a line, p ∈ . 13.2 13.2. Thus the condition for the motion γ to be geodetic means ˙γ dt = o. Let γ(t) = u1(t), u2(t) . Hence we need to put γ(t) = u1(t), u2(t) into the relation .(12.3) and require the right hand side to be zero. I.e. geodetic12 motions are solutions of a system of two differential equations of the 2. order: e13.1e13.1 (1) d2ui dt2 + 2 j,k=1 Γi jk u(t) duj dt duk dt = 0 , i = 1, 2 . This system behaves more or less similarly as one differential equations of teh 2. order. 13.3 13.3. It is well known that a solution of a differential equation of the 2. order is fully determined by its initial value and initial velocity (i.e. the value of its derivative). Analogously, in teh case of the system (13.1) we have Theorem. For every point p ∈ S and every vector A ∈ TpS there exists an interval 0 ∈ IA ⊂ R and a unique geodetic motion γA : IA → S such that γ(0) = p and ˙γ(0) = A. The interval IA generally changes depending on A. 13.4 13.4. It is useful to observe the following property. Lemma. If γ(t) is a geodetic motion then also the motion γ(at + b) is geodetic for each a, b ∈ R, a = 0. Proof. Put γ(t) = γ(at + b). We have dγ dt = dγ dt a, d2γ(t) dt2 = d2γ dt2 a2. Multiplying equations by a2 then also γ(t) = u1(t), u2(t) satisfiese13.1 d2ui dt2 + 2 j,k=1 Γi jk u(t) duj dt duk dt = 0 , i = 1, 2 . 81 This lemma has a natural kinematic interpretation in the plane: If we change parametrization of a linear stadey motion, we obtain again a linear stady motion. 13.5 13.5 Definition. The curve C ⊂ S is called geodetic curve, if there exists such a parametrization γ(t) that γ is a geodetic motion. We briefly talk about geodetics. Geodetic curves in the plane are lines. It follows from 13.3 and 13.4 that Theorem. For each point p ∈ S and each direction in TpS there exists a unique geodetic on S which touches this direction at the point p. 13.6 13.6. It follows form properties of solutions of systems of differential equations of the 2. order (which we shall not discuss here in detail) that Theorem. At each point p ∈ S there exists a neighbourhood Z ⊂ S such that for each two points q1 = q2 in Z there exists a unique geodetic in Z through points q1 and q2. This property is analogous to the fact that each two points in the plane can be connected by a unique line. However, a simple examples shows that the locality assumption in the theorem is essential on surfaces. We shall show below that geodetic curves on the sphere S are great circles. Given any point q1 ∈ S and another point q2 (different from the “opposite pole”), there is a unique great circe throught q1 and q2. However, if q2 is the “opposite pole” to q1 then there are infinitely many great circles throught points q1 and q2. 13.7 13.7. Recall the osculatung plane of a curve is not defined in inflection points. The following theorem provides “outer” characterization of geodesics. Theorem. The curve C ⊂ S is geodesic if and only of its osculating plane at each point contains the normal direction of the surface or is not defined. Proof. The condition ˙γ dt = o means that the vector lies in the normal direction. If d ˙γ dt = o then vectors ˙γ and d ˙γ dt determine the osculating plane which contains the normal direction. If d ˙γ dt = o then this is teh inflection point. In the opposite direction, consider a curve C parametrizad by the arc-length γ(s). Then ˙γ(s), ˙γ(s) = 1. Differentiating the latter, we obtain ˙γ, d ˙γ ds = 0. Je-li d ˙γ ds = o then vectors ˙γ and d ˙γ ds determine the osculating plane and we assume this plane contains the normal line of the surface. Since the vector d ˙γ ds is perpendicular to the vector ˙γ(s), it is parallel with the normal line. Thus ˙γ ds = o. If d ˙γ ds = o then ˙γ ds = o. 82 Example. The great circle C on the sphere S is such circle whose center coincides with the center of the sphere. Its usculating plane coincides with the plane the circle lies in (at every point). Normal lines of the sphere along C lie in the same plane. Hence each great circle is a geodesic. On the other hand, at each point p ∈ S and each direction in TpS, there is a unique great circle. Using the theorem 13.5, other geodesics on S do not exist. 13.8 13.8 Corollary. If C is a geodesic curve and γ(s) its arc-length parametrization then γ(s) is a geodesic motion. Proof. It follows from the theorem 13.7, the osculating plane of the curve contains the normal direction of the surface. The second part of the proof shows that ˙γ ds = o. 13.9 13.9. Let Z ⊂ S is a neighbourhood of the point p ∈ S with the property from the theorem 13.6 Theorem. For each q ∈ Z, q = p, the geodesic through points p and q is the shortes curve in Z which connects points p and q. JS: missing picture Proof. Let us denote by C the geodesic connecting points p and q. Choose a curve ¯C through the point p perpendicular to C. Further, we have a geodetic curve through every point of ¯C perpendicular to ¯C; these curves form a 1-parameter family of parametric curves. We choose its orthogonal trajectories as the second family of parametric curves. Choosing a parametr u1 on C and a parameter u2 on ¯C, we obtain a parametrization of the surface on some neighbourhood U of the point p. We can put p = (0, 0), q = (a, 0), a > 0. Parametrizing curves u2 = c by the arc-length s, the parametrization u1 = s, is a geodetic motion. This it satisfies equations (13). We have du2 ds = 0, d2u2 ds2 = 0, since u2 = c, and further du1 ds = 1 (the parameter is arc-length) and d2u1 ds2 = 0. Putting this into (13.1), we obtain Γ1 11 = 0 , Γ2 11 = 0 . Since the parametric net is orthogonal, we have g12 = (f1, f2) = 0. Using the decomposition 12.(1), we obtain dg11 du1 = ∂(f1, f1) ∂u1 = 2(f1, f11) = 2Γ1 11(f1, f1) = 0 . 83 Differentiating (f1, f2) = 0, we find 0 = ∂(f1, f2) ∂u1 = (f11, f2) + (f1, f12) = Γ2 11(f2, f2) + (f1, f12) = 0 , hence (f1, f12) = 0. Now we obtain ∂g11 ∂u2 = ∂(f1, f1) ∂u2 = 2(f1, f12) = 0 . Thus g11 is a constant k > 0. Consider a curve from p to q given by the parametrization u1 = t, u2 = f(t), f(0) = 0, f(a) = 0. Its length is equal to e13.2e13.2 (2) a 0 k + g22 t, f(t) df dt 2 dt . The length of the geodesic C is a 0 √ k dt = a √ k. Since g22 > 0, the integral (2) is greater or equal to √ ka and the equality holds obly for df dt = 0. Using f(0) = 0, this means that f(t) = 0 for all t. 13.10 13.10 Remark. The oldest approach to the notion of geodesics comes exactly from the property that these are shortest curves connecting two points on the surface. Thus differential equations were derived using the calculus of variation. 13.11 13.11. The curvature κ of a plane curve f(s) satisfies κ = de1 ds , e1 = df ds . An analogy of this property is used to define geodetic curvature of the curve C ⊂ S. Assume that C is parametrized by the arclength γ(s). Then ˙γ = dγ ds is a unit vector. Definition. Geodetic curvature κg of the curve γ(s) on the surface S is defined via the relation κg = ˙γ ds . That is, the geodetic curvature belongs to the inner geometry of the surface. Remark. Also a notion of geodetic torsion of a curve on the surface can be defined but this no more belongs to the inner geometry of the surface. 13.12 13.12. We shall show hot the usual curvature κ and the geodetic curvature κg of a curve C on the surface S are related. 84 Theorem. Let α be the angle between the normal direction of the surface and the osculating plane of the curve C ⊂ S. Then κg = κ sin α. Further, we have κg = 0 in inflection points of the curve C. Proof. In a non-inflection point, the vector d ˙γ ds lies in the osculating plane. JS: missing picture It follows from the picture depictin the section by the plane perpendiculatr to the osculating plane ω of the curve C that ˙γ ds = d ˙γ ds sin α. In inflection points, we have d ˙γ ds = o thus also ˙γ ds = o. 13.13 13.13 Corollary. The curve C on the surface S is geodetic if and only if κg = 0 at all points of C. Proof. This follows directly from theorems 13.7 and 13.12. Lines in a plane are characterized by the property κ = 0. This is one of analogies between lines on a plane and geodetic curves on a surface. 13.14 13.14. Using geodesics, we can extend certain constructions from the euclidean planes to surfaces. The simplest example is the notion of geodetic circles on the surface S. It follows from properties of solutions of systems of 2nd order differential equations that for each point p ∈ S, there is a number rp > 0 such that for each 0 < r < rp and on every geodesic through p there are exactly two points q1 and q2 (one in each direction) such that the length of the arc between p and q1 and of the arc between p and q2 is equal to r. Moreover, it holds that the set K(p, r) of all such points is a curve on S. Definition. The curve K(p, r) is called geodetic circle of the radius r with the center p on the surface S. In the case of the sphere S with the raidus , we shall show that this constraction works generally onlu locally. If r < π then K(p, r) is a usual circle on S which is a curve. For r = π , one moves along geodesics in all directions to the point opposite to p on S. In a sense, putting r = π , the circle K(p, r) “collapses” into a single point. 13.15 13.15. Geodetic circles have constant curvature in the plane and the same is true (as we have just shown) on spheres. This is not true in general, however. Already ellipsoid with axis of different lengths is an example of a surface where geodetic cicles do not have constant geodetic curvature. (We shall mention one more statement in this directions later in 14.7.) 85 14 Surfaces with constant Gaussian curvature The main aim of this chapter is to show a relation between inner geometry of surfaces with noneuclidean geometries. These observation are not only geometrically nice but also played an important role in the history of mathematics. It was mainly understanding of inner geometry which (in the beginning of the 19. century) motiovated leading geometres to believe that noneuclidean geometry forms a mathematical theory in the same sense as euclidean geometry. Proofs in this chapter require either too extensive computations or tools going far beyond theory built so far. These proofs will be omitted — we are interested more in a general view than in a detailed technical matters. 14.1 14.1. We shall investigate local isometries of S with itself. Considering an arbitrary surface, the locallity assumption is so natural that the word “local” will be omitted. Consider a surface of revolution S from 8.15 with the parameter t changed to u. Hence the parametric expression of S is e14.1e14.1 (1) f(u, v) = g(u) cos v, g(u) sin v, h(u) . We found g11 = g 2+h 2, g12 = 0, g22 = g2 in 8.15. The surface of revolution obviously has a 1-parameter system of isometries given by rotations. One can see that also from the 1. fundamental form e14.2e14.2 (2) Φ1 = g11(u) du2 + g22(u) dv2 , g11 = g 2 + h 2 , g22 = g2 . The mapping u = ¯u, v = ¯v + c preserves this form since du = d¯u, dv = d¯v. 14.2 14.2. Assume in the opposite direction that the surface S has a 1-parameter system of isometries such that its trajectories form the system L of curves. Consider its orthogonal trajectories L . Curves form the systemL transform to each other by isometries since an isometry preserves angles. Choose L for coordinate curves u = konst. and L for coordinate curves v = konst. Our isometries than have the form ¯u = u, ¯v = v + c. The form Φ1 is preserved by isometries and also g12 = 0 by orthogonality of coordinate net. Hence e14.3e14.3 (3) Φ1 = A1(u) du2 + A2(u) dv2 , A1 > 0 , A2 > 0 . Change the parameter u to ¯u such that d¯u = √ A1 du, i.e. ¯u = √ A1 du where we keep ¯v = v. Then we have Φ1 = d¯u2 + B(¯u) d¯v2 , B > 0 . 86 This is the 1. fundamental form of a surface of revolution given by graph of the function x = B(z) where we parametrize this curve by arc-length. Summarizing, we have proved Theorem. If the surface S has a 1-parameter system of isometries then it is locally isometric to a surface of revolution. 14.3 14.3. We know that the Gaussian curvature is preserved by isometries. Hence if the surface S has more then one 1-parameter system of isometries then the Gaussian curvature must be constant. We know from 11.19 that a surface is locally isometric to a plane if it has zero Gaussian curvature. However, this holds more generally: two surfaces with the same constant Gaussian curvature are locally isometric, see 14.6 below. The basic example of a surface with zero K is the plane. The basic example of a surface with a constant positive curvature K = 1 r2 is the sphere with radius r, see ??. Now we shall present an example of a surface with a constant negative curvature K = − 1 a2 . 14.4 14.4. We shall need to formula for the Gaussian curvature of the surface of revolution (1). We computed coefficients of 1. and 2. fundamental form in 8.15, e14.4e14.4 (4) K = h11h22 − h2 12 g11g22 − g2 12 = h (g h − h g ) g(g 2 + h 2)2 . Since rotations are isometries, K is independent on the rotation parameter v. 14.5. A plane curve with the parametric expression e14.5e14.5 (5) x = a sin u = g(u) , z = a ln tg u 2 + cos u = h(u) , u ∈ (0, π), u = π 2 is called traktrix with the parameter a > 0, see the picture. (Given JS: Missing picture u = π 2 , we have x = a and lim u→ π 2 h(u) = 0. The point (a, 0) does not lie on the curve, however. This is a sort of singularity.) Rotation around the z-axis yields the surface called pseudosphere. Its Gaussian curvature can be computed using (4). We have g = a cos u, g = −a sin u, h = a 1 tg u 2 cos2 u 2 · 1 2 −sin u = a 1 sin u −sin u , h = a − cos u sin2 u −cos u . 87 Putting this to (4), we obtain K = − 1 a2 . Thus pseudosphere with the parameter a has negative constant Gauassian curvature − 1 a2 . In a sense, this is “opposite” of the sphere with K = 1 r2 which also motivates the terminology 14.6 14.6. We know that isometries of the plane are Euclidean transformations, i.e. plane has 3-parameter system of isometries. Isometries of the sphere S are given by restriction of Euclidean motions in E3 which preserve S. Geometrically, we can easily see that for each pair of points p, ¯p ∈ S and each pair of unit perpendicular vectors e1, e2 ∈ TpS and ¯e1, ¯e2 ∈ T¯pS, there exists a unique isometry g: S → S such that g(p) = ¯p, Tpg(e1) = ¯e1, Tpg(e2) = ¯e2. A similar statement holds for an arbitrary suraface with a constant Gaussian curvature. We shall present this without proof. Theorem (Minding). Let S and ¯S are surfaces with the same constant Gaussian curvature. Then for each pair of points p ∈ S and ¯p ∈ ¯S and each pair of unit perpendicular vectors e1, e2 ∈ TpS and ¯e1, ¯e2 ∈ T¯p ¯S, there exists a unique local isometry g from S to ¯S such that g(p) = ¯p, Tpg(e1) = ¯e1, Tpg(e2) = ¯e2. Hence local isometries of every surface with constant Gaussian curvature form a 3-parameter system of isometries (as in the plane). 14.7 14.7. Consider the geodetic circle K(p, r) on the surface S with a constant Gaussian curvature. It follows from Theorem 14.6 that there is a 1-parameter system of local isometries on S which preserve the point p (in a sense, these are rotations around the point p). Considering small r, this means that the geodetic curvature of the geodetic circle is the same at all points. The case of the sphere was already mentioned 13.15. One can show this holds also globally: Theorem. Geodetic circles on surfaces with constant Gaussian curvature have constant geodetic curvature. 14.8 14.8. Our next tool will be the Gauss-Bonnet theorem. First we shall formulate necessary mathematical terminology. Let C be a curveje with the parametrization f(t), t ∈ I. Definition. Segment U of the curve C corresponding to a closed interval [a, b] ⊂ I is the set f(t), t ∈ [a, b]. 14.9 14.9. Let D ⊂ R2 be an open set. A simple region Ω ⊂ D is an open, convex and bounded subset such that also its closure Ω lies in D and its border ∂Ω is formed by a finite number of segments of curves. JS: missing picture 88 Definition. The subset W ⊂ S is called simple region on the surface S if there exists such a parametrization f : D → E3 of the surface S such that W = f(Ω) where Ω is a simple region in D. Note that one should carefully distinguish between a simple surface and a simple region on a surface. 14.10 14.10. Let h be a function defined on the curve C. Using the given parametrization f : I → E3 of the curve C, this is the function h(t): I → R. Consider the segment U of the curve C which corresponds to the interval [a, b] ⊂ I. Then we define the integral U h ds as e14.6e14.6 (6) U h ds = b a h(t) (f1)2 + (f2)2 + (f3)2 dt . This definition is independent on the choice of parametrization of the curve. We define the integral C h ds using deompocition of the curve C to segments. 14.11 14.11. Let h be a function defined on the surface S. Given a parametrization f : D → E3, this is the function h(u1, u2): D → R. Let Ω ⊂ D be bounded region such that Ω ⊂ D. PWe write W = f(Ω). Recall that we introduced the volume element dV = g11g22 − g2 12 du1 du2 of the surface in ??. The integral W h dV is defined by e14.7e14.7 (7) W h dV = Ω h(u1, u2) g11g22 − g2 12 du1 du2 . Also this definition is independent on the choice of a parametrization of the surface. 14.12 14.12. Now we shall state (without a proof) one of most interesting results of the inner geometry of surfaces. Theorem (Gauss-Bonnet theorem). Let W be a simple region on the surface S whose border is the curve C of the class C2. Let κg be the geodetic curvature of the curve C and K be the Gaussian curvature of the surface S. Then e14.8e14.8 (8) W K dV = 2π − C κg ds . 89 Example. To get a first experience with this theorem, we shall discuss the case of the simple region Ω in the plane whose border is the curve C of the class C2. Its geodetic curvature is the usual curvature and we have K = 0 for the plane. Thus (8) yields e14.9e14.9 (9) C κ ds = 2π . In the case of the circle f(t) = (r cos t, r sin t), t ∈ [0, 2π), this can be verified also by a direct computation. We have κ = 1 r , ds = r2 cos2 t + r2 sin2 t dt = rdt. Thus C κ ds = 2π 0 1 r rdt = 2π . Observe moreover that we have t2 t1 κ ds = t2 − t1 in this case. 14.13 14.13. One can extend the Gauss-Bonnet theorem also to some simple regions whose border is not differentiable because of “vertices”. We shall discuss the case of the curvilinear triangle on the surface S. In the following definition, we understand a triangle in R2 as a closed set including its inner part. Definition. The set W ⊂ S is called curvilinear triangle on the surface S if there exists such local parametrization f : D → E3 of the surface S and such triangle ∆ ⊂ D that W = f(∆). Thus a curvilinear triangle is a simple region on S. Its edges are segments U1, U2, U3 of curves on S. Denote by β1, β2, β3 inner angles of its tangent lines at vertices p1, p2, p3 as on the picture. JS: missing picture Theorem (Generalised Gauss-Bonnet). It holds e14.10e14.10 (10) W K dV = 2π − 3 i=1 βi − 3 i=1 Ui κg ds . Idea of the proof. Let us “smooth” the border of W at the vertex pi using a geodesic circle with Ci with a small radius r as on the picture. JS: missing picture Then one can show that lim r→0 Ci κg ds = βi. One can formulate this also as the idea that as a limit, we have the same situation as in the plane which we discussed at the end of example 14.12. It follows from (8) that be obtain the formulae (10) as a limit. 90 de14.14 14.14 Definition. A curvilinear triangle W is called geodetic triangle on the surface S if all its edges are segments of geodetic curves. This notion is a direct a straightforward modification of the notion of a triangle for the inner geometry of surfaces. We κg = 0 in (10) in this case. This shows Corollary. A geodetic triangle W on the surface S satisfies e14.11e14.11 (11) W K dV = 2π − β1 − β2 − β3 . 14.15 14.15. The following geometrically very interseting statement follows directly from (11): Theorem (Special Gauss-Bonnet). If W is a geodetic triangle on a surface with constant Gauss curvature K of the area V and inner angles α1, α2, α3 then the following holds: e14.12e14.12 (12) α1 + α2 + α3 = π + KV . Proof. PFor K constant, the integral in (11) is equal KV and inner angles are given by the relation βi = π − αi, i = 1, 2, 3. 14.16. Examples. a) Considering developable surfaces, we have K = 014.16 as in the plane. Then (12) yields the well known theorem about the sum of angles in a triangle. α1 + α2 + α3 = π. b) Considering the sphere S with radius r, we have K = 1 r2 . Thus the sum of angles in a geodetic triangle is greater then π. In fact, it follows from (12) that the sum af angles minus π is proportional to the area of the geodetic triangle. It is interesting to realiye that one can compute are of the sphere from (12). Since geodetic curves on S are great circles, one eighth of the sphere is a geodetic triangle with all three right angles. Denote by V its area. Since the Gaussian curvature of the sphere with radius r is 1 r2 , it follows from (12) that 3π 2 = π + 1 r2 V , i.e. V = 1 2πr2. 14.17 14.17. The 5th Euclid axiom states that for a given line p in the plane, there is a unique line nonintersecting p through every point A /∈ p. Nowadays, this is known as the Euclid axiom about parallel lines. This statement is so essentially different from other Euclid axioms that many mathematicians for centuries tried to rpove that the 5th axiom is a consequence of remaining Euclid axioms. However, it was not possible to prove, despite of many (incorrect) attempts, to show that the opposite of the Euclid axiom about parallel lines leads to a contradiction. 91 Thus one of consequences of negation of the 5. axiom, i.e. an assumption of existence at least two lines through the point A nonitersecting p, is the statement that the sum Σof angles in a triangle is smalles than π. Moreover, the difference π−Σ, so called angle defect, is propotional to the area of triangle. Such statement seemed absurd to many mathematicians. The special Gauss-Bonnet theorem shows this case happens in the inner geometry of a surface with negatiove constant Gaussian curvature. Also our theorem 14.6 about the three-parameter system of local isometries on a surface with constant Gaussian curvature corresponds to Euclidean transformations in the classical geometry of plane – with the Euklid axiom about parallel lines or with its negation. Precise constructions of noneuclidean geometries were found in the 2nd half of the 19. century (using essentially projective geometry tools). 14.18. Negation of the 5. axiom leads only to one type of noneuclidean geometries which are called Lobachevsky geometries or hyperbolic geometries. These correspond to surfaces with negative constant Gaussian curvature. However, D. Hilbert shown in the year 1901 that Lobachevsky plane cannot be globaly realized on a surface in E3. (This is the essential explanation why the tractrix in ??, whose rotation yiedls the pseudosphere, has a singular point. An important analogy between properties of surfaces with a posotive and negative constant Gaussian curvature, stated in particular in theorems 14.3, 14.6 and 14.15, motivated to include also elliptic geometries (often related to the name B. Riemann) into noneuclidean geometries. These correspond to the inner geometry of surfaces with positive constant Gaussian curvature. Let us not the notion of Riemannian geometry is usually using in a meaning different from elliptic geometries. In fact, Riemannain geometry denotes far more important theory, which B. Riemann iniciated, and which is nowadays one of more important parts of differential geometry, see e.g. the textbook [5] for a basic information. 92 References 1 [1] J. Bureˇs, K. Hrubk, Diferentiaci´aln geometrie kˇrivek a ploch, Skriptum, UK Praha, 1998. 2 [2] M. P. Do Carmo, Differential geometry of Curves and Surfaces, Prentice-Hall, New Jersey, 1976. 3 [3] M. Doupovec, Diferenci´aln geometrie a tenzorov´y poˇcet, Skriptum, VUT Brno, 1999. 4 [4] W. Klingenberg, A Course in Differential Geometry, Springer-Verlag, 1978. 5 [5] I. Kol´aˇr, ´Uvod do glob´aln anal´yzy, Skritpum, MU Brno, 2003. 6 [6] A. Vanˇzurov´a, Diferenci´aln geometrie kˇrivek a ploch, Skriptum, PU Olomouc, 1996. 93