Skip to main content
Log in

Incidences Between Points and Lines in \({\mathbb {R}}^4\)

  • Published:
Discrete & Computational Geometry Aims and scope Submit manuscript

Abstract

We show that the number of incidences between m distinct points and n distinct lines in \({\mathbb {R}}^4\) is \(O(2^{c\sqrt{\log m}} (m^{2/5}n^{4/5}+m) + m^{1/2}n^{1/2}q^{1/4} + m^{2/3}n^{1/3}s^{1/3} + n)\), for a suitable absolute constant c, provided that no 2-plane contains more than s input lines, and no hyperplane or quadric contains more than q lines. The bound holds without the factor \(2^{c\sqrt{\log m}}\) when \(m \le n^{6/7}\) or \(m \ge n^{5/3}\). Except for the factor \(2^{c\sqrt{\log m}}\), the bound is tight in the worst case.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. The additional requirement in [15], that no regulus contains too many lines, is not needed for the incidence bound given below.

  2. In this overview we assume some familiarity of the reader with the new “polynomial method” of Guth and Katz, and with subsequent applications thereof. Otherwise, the overview can be skipped on first reading.

  3. That is, every point \(p\in Z(f)\) is incident to a line that is fully contained in Z(f); see Salmon [8, 15, 24, 32, 37] for definitions.

  4. This is not quite the case: Guth and Katz also require that no regulus contains more than s (actually, \(\sqrt{n}\)) lines, but this is made to bound the number of points incident to just two lines, and is not needed for the incidence bound in Theorem 1.1.

  5. The meaning of this statement is that the assertion holds for the fiber at any point outside some lower-dimensional exceptional subvariety.

  6. The Zariski closure of a set Y is the intersection of all varieties X that contain Y. Y is Zariski closed if it is equal to its closure (and is therefore a variety), and is Zariski open if its complement is Zariski closed. See [18] for further details.

  7. For example, for the surface in \({\mathbb {R}}^3\) defined by the zero set of \(f=x+y+z+x^3\), the point \(0=(0,0,0)\in Z(f)\) is flat (because the second order Taylor expansion of f near 0 is the plane \(x+y+z=0\)), but is not linearly flat, since there is no line incident to 0 and contained in Z(f).

  8. This property holds for \({\mathbb {C}}\) but not for \({\mathbb {R}}\).

  9. When \(m\le \sqrt{n}\) (or when \(n \le \sqrt{m}\)), an immediate application of the Szemerédi–Trotter theorem yields the linear bound \(O(m+n)\).

  10. Note that in general the bounds \(O(D^2)\) and O(D) are not necessarily smaller than their respective original counterparts q and s. Nevertheless, they uniformly depend on m and n in a way that makes them fit the induction process, whereas the parameters q and s, over which we have no control, do not.

  11. As the calculations worked out above indicate, the bounds in Proposition 3.3 will be within the bound (5) when m is sufficiently small (below \(n^{4/3}\)) or sufficiently large (above \(n^{5/3}\)). For such values of m we can bypass the induction process, and obtain the desired bounds directly, in a single step. See a more detailed description towards the end of this section.

  12. Similar to the definition in Sect. 2.4 for the case of lines, it suffices to require this property for every point in some Zariski-open subset of X. Here too one can show that the two definitions are equivalent. See also the companion paper [37, Lem. 11].

  13. This rather minuscule value of E is only needed when \(m \approx n^{4/3}\); for smaller values of m, much larger values of E can be chosen.

References

  1. Basu, S., Sombra, M.: Polynomial partitioning on varieties of codimension two and point-hypersurface incidences in four dimensions. Discrete Comput. Geom. 55(1), 158–184 (2016). Also in arXiv:1406.2144

    Article  MathSciNet  MATH  Google Scholar 

  2. Beauville, A.: Complex Algebraic Surfaces, vol. 34. Cambridge University Press, Cambridge (1996)

    Book  MATH  Google Scholar 

  3. Clarkson, K., Edelsbrunner, H., Guibas, L., Sharir, M., Welzl, E.: Combinatorial complexity bounds for arrangements of curves and spheres. Discrete Comput. Geom. 5, 99–160 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  4. Cox, D., Little, J., O’Shea, D.: Using Algebraic Geometry. Springer, Heidelberg (2005)

    MATH  Google Scholar 

  5. Cox, D., Little, J., O’Shea, D.: Ideals, Varieties, and Algorithms: An Introduction to Computational Algebraic Geometry and Commutative Algebra. Springer, Heidelberg (2007)

    Book  MATH  Google Scholar 

  6. Dvir, Z.: On the size of Kakeya sets in finite fields. J. Am. Math. Soc. 22, 1093–1097 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  7. Dvir, Z., Gopi, S.: On the number of rich lines in truly high dimensional sets. In: Proceedings of 30th Annual ACM Symposium on Computational Geometry, pp. 584–598 (2015)

  8. Edge, W.L.: The Theory of Ruled Surfaces. Cambridge University Press, Cambridge (2011)

    MATH  Google Scholar 

  9. Elekes, G.: Sums versus products in number theory, algebra and Erdős geometry—a survey. Paul Erdős and His Mathematics II. Bolyai Mathematical Society Studies, vol. 11, pp. 241–290. Bolyai Mathematical Society, Budapest (2002)

    Google Scholar 

  10. Elekes, G., Kaplan, H., Sharir, M.: On lines, joints, and incidences in three dimensions. J. Comb. Theory, Ser. A 118, 962–977 (2011). Also in arXiv:0905.1583

    Article  MathSciNet  MATH  Google Scholar 

  11. Fuchs, D., Tabachnikov, S.: Mathematical Omnibus: Thirty Lectures on Classic Mathematics. American Mathematical Society, Providence, RI (2007)

    Book  MATH  Google Scholar 

  12. Fulton, W.: Introduction to Intersection Theory in Algebraic Geometry, Expository Lectures from the CBMS Regional Conference Held at George Mason University, June 27–July 1, 1983, Vol. 54. AMS Bookstore (1984)

  13. Guth, L.: Distinct distance estimates and low-degree polynomial partitioning. Discrete Comput. Geom. 48, 1–17 (2014). Also in arXiv:1404.2321

    Google Scholar 

  14. Guth, L., Katz, N.H.: Algebraic methods in discrete analogs of the Kakeya problem. Adv. Math. 225, 2828–2839 (2010). Also in arXiv:0812.1043v1

    Article  MathSciNet  MATH  Google Scholar 

  15. Guth, L., Katz, N.H.: On the Erdős distinct distances problem in the plane. Ann. Math. 181, 155–190 (2015). Also in arXiv:1011.4105

    Article  MathSciNet  MATH  Google Scholar 

  16. Hablicsek, M., Scherr, Z.: On the number of rich lines in high dimensional real vector spaces. Discrete Comput. Geom. 55(1), 1–8 (2014). Also in arXiv:1412.7025

    MathSciNet  MATH  Google Scholar 

  17. Harris, J.: Algebraic Geometry: A First Course, vol. 133. Springer, New York (1992)

    MATH  Google Scholar 

  18. Hartshorne, R.: Algebraic Geometry. Springer, New York (1983)

    MATH  Google Scholar 

  19. Ivey, T.A., Landsberg, J.M.: Cartan for Beginners: Differential Geometry via Moving Frames and Exterior Differential Systems, Graduate Studies in Mathematics, vol. 61. American Mathematical Society, Providence, RI (2003)

    MATH  Google Scholar 

  20. Kaplan, H., Sharir, M., Shustin, E.: On lines and joints. Discrete Comput. Geom. 44, 838–843 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  21. Kaplan, H., Matoušek, J., Safernová, Z., Sharir, M.: Unit distances in three dimensions. Comb. Prob. Comput. 21, 597–610 (2012). Also in arXiv:1107.1077

    Article  MathSciNet  MATH  Google Scholar 

  22. Kaplan, H., Matoušek, J., Sharir, M.: Simple proofs of classical theorems in discrete geometry via the Guth–Katz polynomial partitioning technique. Discrete Comput. Geom. 48, 499–517 (2012). Also in arXiv:1102.5391

    Article  MathSciNet  MATH  Google Scholar 

  23. Katz, N.H.: The flecnode polynomial: a central object in incidence geometry, in arXiv:1404.3412

  24. Kollár, J.: Szemerédi–Trotter-type theorems in dimension 3. Adv. Math. 271, 30–61 (2015). Also in arXiv:1405.2243

    Article  MathSciNet  MATH  Google Scholar 

  25. Landsberg, J.M.: Is a linear space contained in a submanifold? On the number of derivatives needed to tell. J. Reine Angew. Math. 508, 53–60 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  26. Mezzetti, E., Portelli, D.: On Threefolds Covered by Lines. Abhandlungen aus dem Mathematischen Seminar der Universität Hamburg, vol. 70, no. 1. Springer, Heidelberg (2000)

    MATH  Google Scholar 

  27. Pach, J., Sharir, M.: Geometric incidences. In: Pach, J. (ed.) Towards a Theory of Geometric Graphs. Contemporary Mathematics, vol. 342, pp. 185–223. American Mathematical Society, Providence, RI (2004)

    Chapter  Google Scholar 

  28. Pressley, A.: Elementary Differential Geometry. Springer Undergraduate Mathematics Series. Springer, London (2001)

    Book  MATH  Google Scholar 

  29. Quilodrán, R.: The joints problem in \(R^n\). SIAM J. Discrete Math. 23(4), 2211–2213 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  30. Richelson, S.: Classifying Varieties with Many Lines, a senior thesis, Harvard University (2008). http://www.math.harvard.edu/theses/senior/richelson/richelson.pdf

  31. Rogora, E.: Varieties with many lines. Manuscr. Math. 82(1), 207–226 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  32. Salmon, G.: A Treatise on the Analytic Geometry of Three Dimensions, vol. 2, 5th edn. Hodges, Figgis and co. Ltd., Dublin (1915)

    MATH  Google Scholar 

  33. Segre, B.: Sulle \(V_n\) contenenti più di \(\infty ^{n-k}S_k\), Nota I e II. Rend. Accad. Naz. Lincei 5(193–197), 275–280 (1948)

    MathSciNet  MATH  Google Scholar 

  34. Severi, F.: Intorno ai punti doppi impropri etc. Rend. Cir. Math. Palermo 15(10), 33–51 (1901)

    Article  MATH  Google Scholar 

  35. Sharir, M., Solomon, N.: Incidences between points and lines in \(\mathbb{R}^4\). In: Proceedings of 30th Annual Symposium on Computational Geometry, pp. 189–197 (2014)

  36. Sharir, M., Solomon, N.: Incidences between points and lines in three dimensions, Intuitive Geometry. Also In: Pach, J. (ed.) Proceedings of 31st Annual Symposium on Computational Geometry. Also in arXiv:1501.02544 (2015)

  37. Sharir, M., Solomon, N.: Incidences between points and lines on a two-dimensional variety, in arXiv:1502.01670

  38. Sharir, M., Sheffer, A., Zahl, J.: Improved bounds for incidences between points and circles. Comb. Prob. Comput. 24(03), 490–520 (2015). Also in arXiv:1208.0053

    Article  MathSciNet  MATH  Google Scholar 

  39. Solomon, N., Zhang, R., Solomon,: Highly incidental patterns on a quadratic hypersurface in \(\mathbb{R}^4\), in arXiv:1601.01817

  40. Solymosi, J., Tao, T.: An incidence theorem in higher dimensions. Discrete Comput. Geom. 48, 255–280 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  41. Székely, L.: Crossing numbers and hard Erdős problems in discrete geometry. Combt. Prob. Comput. 6, 353–358 (1997)

    Article  MATH  Google Scholar 

  42. Szemerédi, E., Trotter, W.T.: Extremal problems in discrete geometry. Combinatorica 3, 381–392 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  43. Tao, T.: From rotating needles to stability of waves: emerging connections between combinatorics, analysis, and PDE. Notices Am. Math. Soc. 48(3), 294–303 (2001)

    MathSciNet  MATH  Google Scholar 

  44. Tao, T.: The Cayley–Salmon theorem via classical differential. https://terrytao.wordpress.com/2014/03/28/the-cayley-salmon-theorem-via-classical-differential-geometry/

  45. van der Waerden, B.L., Artin, E., Noether, E.: Modern Algebra, vol. 2. Springer, Heidelberg (1966)

    Google Scholar 

  46. Warren, H.E.: Lower bound for approximation by nonlinear manifolds. Trans. Am. Math. Soc. 133, 167–178 (1968)

    Article  MathSciNet  MATH  Google Scholar 

  47. Zahl, J.: An improved bound on the number of point-surface incidences in three dimensions. Contrib. Discrete Math. 8(1), 100–121 (2013). Also in arXiv:1104.4987

    MathSciNet  MATH  Google Scholar 

  48. Zahl, J.: A Szemerédi–Trotter type theorem in \(\mathbb{R}^4\). Discrete Comput. Geom. 54(3), 513–572 (2015). Also in arXiv:1203.4600

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

Work on this paper by Noam Solomon and Micha Sharir was supported by Grant 892/13 from the Israel Science Foundation. Work by Micha Sharir was also supported by Grant 2012/229 from the U.S.–Israel Binational Science Foundation, by the Israeli Centers of Research Excellence (I-CORE) program (Center No. 4/11), and by the Hermann Minkowski-MINERVA Center for Geometry at Tel Aviv University. We would like to thank several people whose advice, comments and guidance have helped us a lot in our work on the paper. They are János Kollár, Martin Sombra, Aise J. de Jong, and Saugata Basu. We also thank the anonymous referees for their helpful comments on the paper. In addition, as noted, part of the work on the paper was carried out during the special semester on Algebraic Techniques for Combinatorial and Computational Geometry, held at the Institute for Pure and Applied Mathematics at UCLA, in the Spring of 2014. We are grateful for the pleasant working environment provided by IPAM, and for the helpful interaction with additional colleagues, including Larry Guth, Nets Hawk Katz, Terry Tao, Jordan Ellenberg, and many others.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Noam Solomon.

Additional information

Editor in Charge: János Pach

Part of this research was performed while the authors were visiting the Institute for Pure and Applied Mathematics (IPAM), which is supported by the National Science Foundation. An earlier version of this study appears in Proc. 30th Annu. ACM Sympos. Comput. Geom., 2014, 189–197, and the present version is also available in arXiv:1411.0777v1.

Appendix: Severi’s Theorem

Appendix: Severi’s Theorem

In this appendix we sketch a proof of Severi’s theorem (Theorem 3.11).

First, recall from Sect. 2.1 that a real (resp., complex) surface X is ruled by real (resp., complex) lines if every point \(p\in X\) in a Zariski open dense set is incident to a real (resp., complex) line that is fully contained in X. This definition has been used in several recent works (see, e.g.,  [15]); this is a slightly weaker condition than the classical condition that requires that every point of X be incident to a line contained in X. Nevertheless, as we show next, the two are equivalent.

Lemma 6.1

Let \(f \in {\mathbb {R}}[x,y,z]\,(\)resp., \(f \in {\mathbb {R}}[x,y,z,w])\) be an irreducible polynomial such that there exists a Zariski open dense set \(U\subseteq Z(f)\), so that each point in the set is incident to a line, fully contained in Z(f). Then \(\mathsf {FL}_{f}\ (\)resp., \(\mathsf {FL}_{f}^4)\) vanishes identically on Z(f), and Z(f) is ruled by lines.

Proof

By assumption and definition, \(\mathsf {FL}_{f}\) (resp., \(\mathsf {FL}_{f}^4\)) vanishes on U. If it vanishes on Z(f), Theorem 2.11 implies that Z(f) is ruled. Otherwise, \(Z(f, \mathsf {FL}_{f})\) (resp., \(Z(f,\mathsf {FL}_{f}^4)\)) is properly contained in Z(f) and contains U. Since Z(f) is irreducible, this latter variety must be of dimension at most 1 (resp., 2). On the other hand, \(Z(f, \mathsf {FL}_{f})\) (resp., \(Z(f,\mathsf {FL}_{f}^4)\)) is Zariski closed set (by definition of the Zariski topology) and therefore contains its Zariski closure. As U is Zariski dense, its Zariski closure is Z(f). \(\square \)

Remark

In Sharir and Solomon [37], we have proved the same statement without using the Flecnode polynomial.

This phenomenon generalizes to k-flats instead of lines (and the proof translates verbatim).

Lemma 6.2

Let V be an irreducible variety for which there exists a Zariski open subset \(U\subseteq V\) with the property that each point \(p\in U\) is incident to a k-flat that is fully contained in V. Then this property holds for every point of V.

We now proceed to sketch a proof of Severi’s theorem. For convenience, we repeat its statement.

Theorem 3.11

(Severi’s Theorem [34]) Let \(X \subset {\mathbb {P}}^d({\mathbb {C}})\) be a k-dimensional irreducible variety, and let \(\Sigma _0\) be a component of maximal dimension of F(X), such that the lines of \(\Sigma _0\) cover X. Then the following holds.

  1. 1.

    If \(\dim (\Sigma _0) = 2k - 2\), then X is a copy of \(\mathbb {P}^k({\mathbb {C}})\,(\)that is, a complex projective k-flat).

  2. 2.

    If \(\dim (\Sigma _0) = 2k -3\), then either X is a quadric, or X is ruled by copies of \({\mathbb {P}}^{k-1}({\mathbb {C}})\), i.e., every point \(p\in X\) is incident to a copy of \({\mathbb {P}}^{k-1}({\mathbb {C}})\) that is fully contained in X.

We sketch a proof in the case \(k=3,d=4\), under the simplifying assumption that for any non-singular \(x\in X,\Sigma _{0,x}\) is infinite; this assumption holds in our application of the theorem (by the informal dimensionality argument mentioned in the paper, it holds “on average” in general for these parameters). Our proof is based on a sketch provided by A. J. de Jong, via private communication, and we are very grateful for his assistance.

Sketch of Proof

For \(x \in X\), we recall that \(\Xi _{0,x}\) denotes the cone of lines (i.e., union of lines) of \(\Sigma _{0,x}\) The proof consists of the following steps.

  1. (1)

    Assume first that \(\dim (\Sigma _0)=2k-2=4\). Then there exists some non-singular point \(x_0\in X\) with \(\dim (\Sigma _{0,x_0})=2\). Indeed, if, for all non-singular points \(x\in X,\dim (\Sigma _{0,x}) \le 1\), then \(\dim (\Sigma _0) < 4\) (see the analysis in Theorem 3.9, and the preceding analysis), contradicting the assumption in this case. By an argument that has already been sketched earlier, this implies that \(\dim (\Xi _{0, x_0})=3\), i.e., the cone of lines in \(\Sigma _{0,x_0}\) through \(x_0\) is three-dimensional, and therefore \(X=\Xi _{0,x_0}\). As \(x_0\) is non-singular, it follows that X must be a hyperplane, as claimed.

  2. (2)

    Consider next the case where \(\dim (\Sigma _0)=2k-3=3\), and for any non-singular point \(x\in X,\Sigma _{0,x}\) is 1-dimensional (as just argued, if \(\Sigma _{0,x}\) is two-dimensional for some non-singular \(x \in X\), then X is a hyperplane). In other words, \(\Sigma _{0,x}\), parameterized by the direction of its lines, is a curve in \({\mathbb {P}} T_x X \cong {\mathbb {P}}^2({\mathbb {C}})\); put \(e_x\) for its degree. If \(e_x = 1\), then \(\Xi _{0,x}\) contains a 2-flat.

We next define a “plane-flecnode polynomial system” associated with X, that expresses, for a point \(x\in X\), the existence of a 2-flat H, such that H osculates to X to order 3 at x. Since X is a hypersurface, we can write \(X=Z(f)\), for a suitable 4-variate polynomial f (see Sect. 2), and assume that f is irreducible (as X is irreducible).

We represent a 2-flat through the origin in \({\mathbb {C}}^4\) (ignoring the lower-dimensional family of 2-flats that cannot be represented in this manner) as

$$\begin{aligned} H_{v_0,v_1,v_2,v_3}:= \{(x,y,z,w) \mid z=v_0x+v_1y, \ w=v_2x+v_3y\}, \end{aligned}$$
(46)

for \(v_0,v_1,v_2,v_3 \in {\mathbb {C}}\). The 2-flat \(H_{v_0,v_1,v_2,v_3}\) is said to osculate to \(X=Z(f)\) to order k at p, if the Taylor expansion of f at p along H satisfies

$$\begin{aligned} f(p+(x,y,v_0x+v_1y, v_2x+v_3y))=O(x^{k+1}+y^{k+1}). \end{aligned}$$
(47)

This translates into a system of homogeneous polynomial equations in \(v_0,v_1,v_2,v_3\), involving the partial derivatives of f up to order k. Specializing to the case \(k=3\), the plane-flecnode polynomial system, \(\mathsf {PFL}_{f}\), associated with f, is obtained by eliminating \(v_0,v_1,v_2,v_3\) from these equations (for osculation up to order 3). This is the multipolynomial resultant system of the polynomials defining these equations up to order 3, with respect to \(v_0,v_1,v_2,v_3\) (see Van der Waerden [45, Chap. XI] for details).

Another theorem of Landsberg [25, Thm. 1] states that, if, for every \(g\in \mathsf {PFL}_{f},g\) vanishes identically on X, then X is ruled by 2-flats, which finishes the proof in this case.

Therefore, we may assume that \(X\cap Z(\mathsf {PFL}_{f})\) is a Zariski closed proper subset of X. By definition of \(\mathsf {PFL}_{f}\), it follows that for every non-singular point \(x\in X \setminus Z(\mathsf {PFL}_{f})\) (namely, outside the Zariski closed set \(Z(\mathsf {PFL}_{f})\)), we have \(e_x>1\). Indeed, if \(e_x=1\), then, as observed above, there is a 2-flat incident to x, and fully contained in X, implying that for every \(g \in \mathsf {PFL}_{f},g(x)=0\), contradicting the assumption that \(x \in X \setminus Z(\mathsf {PFL}_{f})\).

For a generic hyperplane H in \({\mathbb {P}}^4({\mathbb {C}})\), which is not contained in \(Z(\mathsf {PFL}_{f})\), put \(S_H := X \cap H\). As observed above, \(X \cap Z(\mathsf {PFL}_{f})\) is properly contained in X, which in turn implies that, for a generic hyperplane H in \({\mathbb {P}}^4({\mathbb {C}}),S_H\) is not fully contained in \(Z(\mathsf {PFL}_{f})\). Indeed, let g be a polynomial in \(\mathsf {PFL}_{f}\) that does not vanish identically on X. Then \(X \cap Z(g) = Z(f,g)\) is strictly contained in \(X=Z(f)\), and since Z(f) is irreducible, it follows that Z(fg) is two-dimensional. Therefore, for a generic hyperplane \(H,X \cap Z(\mathsf {PFL}_{f}) \cap H\) is contained in the one-dimensional variety \(Z(f,g)\cap H\), and thus cannot contain the two-dimensional variety \(S_H\).

Let \(x \in X\) be a non-singular point, and let H be a hyperplane in \({\mathbb {P}}^4({\mathbb {C}})\), which is incident to X and not contained in \(Z(\mathsf {PFL}_{f})\). We claim that for a generic H, there are \(e_x\) distinct lines that are incident to x and fully contained in \(S_H\). Indeed, the intersection of the hyperplane H with \(T_x X\) is a 2-flat in \(T_x X\) containing x. Taking its projectivization (where the point x is regarded as 0), namely, \({\mathbb {P}} T_x X \cong {\mathbb {P}}^2\), the (generic) 2-flat \(T_x X \cap H\) becomes a (generic) line. The degree of \(\Sigma _{0,x} \subset {\mathbb {P}} T_x X\) is \(e_x\). Therefore, the intersection of \(\Sigma _x\) with a line in \({\mathbb {P}} T_x X\cong {\mathbb {P}}^2\) consists of \(e_x\) points, which are distinct since the line is generic. Therefore, its intersection with \(\Sigma _{0,x}\) consists of \(e_x\) distinct points. These \(e_x\) distinct (projective) points represent \(e_x\) distinct lines, incident to x and fully contained in \(X\cap H=S_H\), as claimed.

We say that a pair (xH), where H is a hyperplane in \({\mathbb {P}}^4({\mathbb {C}})\) and \(x \in S_H\), is adequate if there are \(e_x\) distinct lines incident to x that are fully contained in \(S_H\). Since a generic point x is non-singular, the previous paragraph implies that a generic pair (xH) is adequate. Therefore, by changing the order of quantifiers, fixing a generic hyperplane H, a generic point \(x \in S_H\) is such that the pair (xH) is adequate.

By Bertini’s Theorem (see, e.g., Harris [17, Thm. 17.16]), the irreducibility of X implies that for a generic hyperplane H, the surface \(S_H\) is an irreducible surface in \(H\cong \mathbb {P}^3({\mathbb {C}})\). For a generic point \(x \in S_H\), that is, outside an algebraic curve \({\mathcal {C}}_H\) in \(S_H\), the pair (xH) is adequate. Therefore, there are \(e_x\) distinct lines that are incident to x and fully contained in \(S_H\), which, by Lemma 6.1, implies that \(S_H\) is a ruled surface. Moreover, for any \(x \in S_H\setminus Z(\mathsf {PFL}_{f})\), we have \(e_x>1\). As observed above, \(\mathsf {PFL}_{f}\) does not vanish identically on \(S_H\), implying that \(Z(\mathsf {PFL}_{f})\cap S_H\) is a Zariski closed proper subset of \(S_H\), i.e., an algebraic curve contained in \(S_H\). Adding this curve to \({\mathcal {C}}_H\), it follows that outside this algebraic curve, each point of \(S_H\) is incident to at least two lines fully contained in \(S_H\). By Sharir and Solomon [37, Lem. 9], this implies that \(S_H\) is either a 2-flat or a regulus. If X is of degree greater than two, then, for a generic hyperplane \(H,S_H\) is a (two-dimensional) surface of degree greater than two. Therefore, X must be of degree at most two, namely, X is either a hyperplane or a quadric. If X is a hyperplane, then \(\Sigma _{0}\) is four-dimensional, contrary to the present assumption, so finally, we deduce that X is a quadric, and the proof is complete. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sharir, M., Solomon, N. Incidences Between Points and Lines in \({\mathbb {R}}^4\) . Discrete Comput Geom 57, 702–756 (2017). https://doi.org/10.1007/s00454-016-9822-2

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00454-016-9822-2

Keywords

Navigation