Ebooks File Surveys in Geometry I Athanase Papadopoulos Editor All Chapters

Download as pdf or txt
Download as pdf or txt
You are on page 1of 49

Full download ebooks at https://ebookmeta.

com

Surveys in Geometry I Athanase Papadopoulos


Editor

For dowload this book click link below


https://ebookmeta.com/product/surveys-in-geometry-i-
athanase-papadopoulos-editor/

OR CLICK BUTTON

DOWLOAD NOW
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Surveys in Geometry I Athanase Papadopoulos

https://ebookmeta.com/product/surveys-in-geometry-i-athanase-
papadopoulos/

Surveys in Combinatorics 2022 Anthony Nixon

https://ebookmeta.com/product/surveys-in-
combinatorics-2022-anthony-nixon/

Surveys in Combinatorics 2021 Konrad K. Dabrowski

https://ebookmeta.com/product/surveys-in-
combinatorics-2021-konrad-k-dabrowski/

Lectures on the Geometry of Manifolds Third Edition


Liviu I Nicolaescu

https://ebookmeta.com/product/lectures-on-the-geometry-of-
manifolds-third-edition-liviu-i-nicolaescu/
Extended Abstracts GEOMVAP 2019 Geometry Topology
Algebra and Applications Women in Geometry and Topology
Trends in Mathematics Maria Alberich-Carramiñana
(Editor)
https://ebookmeta.com/product/extended-abstracts-
geomvap-2019-geometry-topology-algebra-and-applications-women-in-
geometry-and-topology-trends-in-mathematics-maria-alberich-
carraminana-editor/

Arakelov Geometry and Diophantine Applications Lecture


Notes in Mathematics Emmanuel Peyre (Editor)

https://ebookmeta.com/product/arakelov-geometry-and-diophantine-
applications-lecture-notes-in-mathematics-emmanuel-peyre-editor/

Vagueness in the Exact Sciences 1st Edition Apostolos


Syropoulos Basil K Papadopoulos

https://ebookmeta.com/product/vagueness-in-the-exact-
sciences-1st-edition-apostolos-syropoulos-basil-k-papadopoulos/

Political Geometry: Rethinking Redistricting in the US


with Math, Law, and Everything In Between Moon Duchin
(Editor)

https://ebookmeta.com/product/political-geometry-rethinking-
redistricting-in-the-us-with-math-law-and-everything-in-between-
moon-duchin-editor/

Free Resolutions in Commutative Algebra and Algebraic


Geometry Research Notes in Mathematics 1st Edition
David Eisenbud (Editor)

https://ebookmeta.com/product/free-resolutions-in-commutative-
algebra-and-algebraic-geometry-research-notes-in-mathematics-1st-
edition-david-eisenbud-editor/
Athanase Papadopoulos Editor

Surveys in
Geometry I
Surveys in Geometry I
Athanase Papadopoulos
Editor

Surveys in Geometry I
Editor
Athanase Papadopoulos
Institut de Recherche Mathématique
Avancée
Université de Strasbourg et CNRS
Strasbourg, France

ISBN 978-3-030-86694-5 ISBN 978-3-030-86695-2 (eBook)


https://doi.org/10.1007/978-3-030-86695-2

Mathematics Subject Classification: 30F10, 30F60, 32G15, 53C70, 51K05, 53A35, 57M60, 52B60,
30F10, 30F60, 32G15, 53C70, 51K05, 53A35, 57M60, 52B60

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland
AG 2022
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG.
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

This is the first of a two-volume set of surveys on geometry in the broad sense
(including group actions and topology). The surveys vary in their scope and degree
of difficulty, but they all represent current research trends.
In planning this book, I naturally wanted to promote topics that I personally
like. Some chapters involve only classical mathematics (e.g., geometry of finite-
dimensional vector spaces and spherical geometry), while others are concerned with
more recent topics (e.g., Gromov-hyperbolic spaces and Teichmüller spaces), but
my personal feeling is that in the end there is little difference between classical
and modern, or between elementary and advanced mathematics; all the ideas, old
and new, are interrelated, and they form one single subject, geometry. Some of the
surveys in this volume are based on lectures that were given by their authors to
students who are at a middle-advanced level. In particular, three surveys consist of
polished notes of lectures given at a CIMPA thematic school that I co-organized
with Bankteshwar Tiwari in 2019, at the Banaras Hindu University in Varanasi.
Other notes associated with lectures delivered at the same school will appear in the
second volume.
When I asked the authors to write a survey for this collection, I knew what to
expect in terms of content, but I emphasized the fact that the goal is to give the
reader a real introduction to the subject, in an attractive way. Most of the authors
succeeded in this, and I take this opportunity to thank them all for their contribution.
My thanks also go to Elena Griniari, from Springer, for her support.

Strasbourg, France Athanase Papadopoulos


September, 2021

v
Contents

1 Introduction .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 1
Athanase Papadopoulos
2 Spherical Geometry—A Survey on Width and Thickness of
Convex Bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 7
Marek Lassak
3 Minkowski Geometry—Some Concepts and Recent
Developments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 49
Vitor Balestro and Horst Martini
4 Orthogonality Types in Normed Linear Spaces . . . . .. . . . . . . . . . . . . . . . . . . . 97
Javier Alonso, Horst Martini, and Senlin Wu
5 Convex Bodies: Mixed Volumes and Inequalities . . .. . . . . . . . . . . . . . . . . . . . 171
Ivan Izmestiev
6 Compactness and Finiteness Results for Gromov-Hyperbolic
Spaces . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 205
Gérard Besson and Gilles Courtois
7 All 4-Dimensional Smooth Schoenflies Balls Are
Geometrically Simply-Connected .. . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 269
Valentin Poénaru
8 Classical Differential Topology and Non-commutative Geometry . . . . 309
Valentin Poénaru
9 A Short Introduction to Translation Surfaces, Veech
Surfaces, and Teichmüller Dynamics . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 343
Daniel Massart

vii
viii Contents

10 Teichmüller Spaces and the Rigidity of Mapping Class


Group Actions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 389
Ken’ichi Ohshika
11 Holomorphic G-Structures and Foliated Cartan Geometries
on Compact Complex Manifolds . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 417
Indranil Biswas and Sorin Dumitrescu

Index . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 463
Contributors

Javier Alonso Instituto de Matemáticas de la UEx (IMUEx), Edificio Carlos


Benítez, Badajoz, Spain
Vitor Balestro Instituto de Matemática e Estatística, Universidade Federal Flumi-
nense, Niterói, Brazil
Gérard Besson Institut Fourier, Université de Grenoble, Saint Martin d’Hères,
France
Indranil Biswas School of Mathematics, Tata Institute of Fundamental Research,
Mumbai, India
Gilles Courtois Institut de Mathématiques de Jussieu – Paris Rive Gauche (IMJ-
PRG), Paris, France
Sorin Dumitrescu Université Côte d’Azur, CNRS, LJAD, Nice, France
Ivan Izmestiev TU Wien, Institute of Discrete Mathematics and Geometry,
Vienna, Austria
Marek Lassak Institute of Mathematics and Physics, University of Science and
Technology, Bydgoszcz, Poland
Daniel Massart Institut Montpelliérain Alexander Grothendieck, CNRS, Univer-
sité de Montpellier, Montpellier, France
Horst Martini Fakultät für Mathematik, Technische Universität Chemnitz, Chem-
nitz, Germany
Ken’ichi Ohshika Department of Mathematics, Faculty of Science, Gakushuin
University, Toshima-ku, Tokyo, Japan
Athanase Papadopoulos Université de Strasbourg and CNRS, Strasbourg, France
Valentin Poénaru Université de Paris-Sud, Mathématiques, Orsay, France
Senlin Wu North University of China, Taiyuan, China

ix
About the Editor

Athanase Papadopoulos (born 1957) is Directeur de Recherche at the French Cen-


tre National de la Recherche Scientifique. His main fields of interest are geometry
and topology, history and philosophy of mathematics, and mathematics and music.
He has held visiting positions at the Institute for Advanced Study, Princeton (1984–
1985 and 1993–1994), USC (1998–1999), CUNY (Ada Peluso Professor, 2014),
Brown University (Distinguished Visiting Professor, 2017), Tsinghua University,
Beijing (2018), and Lamé Chair of the State University of Saint Petersburg (2019)
and has made several month visits to the Max Planck Institute for Mathematics
(Bonn), Erwin Schrödinger Institute (Vienna), Graduate Center of CUNY (New
York), Tata Institute (Bombay), Galatasaray University (Istanbul), University of
Florence (Italy), Fudan University (Shanghai), Gakushuin University (Tokyo), and
Presidency University (Calcutta). He is the author of more than 200 published
articles and 35 monographs and edited books.

xi
Chapter 1
Introduction

Athanase Papadopoulos

Abstract This introductory chapter contains a general description of the subject


matter and a detailed outline of the content of the book.

Keywords Spherical geometry · Minkowski geometry · Finiteness theorems in


Gromov-hyperbolic space · Convex geometry · 4-Manifolds · Differential
topology · Smooth Schoenflies ball · Geometrically simple connectivity ·
Translation surface · Veech surface · Teichmüller space · Mapping class group ·
Holomorphic structure · Foliated Cartan geometry

The present volume of surveys covers a large spectrum of current research topics
in geometry in a broad sense, including spherical geometry, infinitesimal geometry
(Riemannian and Finsler), metric spaces à la Gromov, Busemann spaces, convexity,
singular flat structures on surfaces and their dynamics, Teichmüller spaces, Cartan
geometries and generalizations, and the topology of 4-manifolds.
The first survey is concerned with the analogue in spherical geometry of a theory
that was previously developed in the Euclidean setting. Working on such a subject
follows the tradition of a series of efforts made by a number of mathematicians to
adapt to a non-Euclidean setting notions and results that were known previously
in Euclidean geometry. Among these authors, let me mention Leonhard Euler, who
published a series of memoirs in which he presented theorems in spherical geometry
that are analogues of results (some of which are classical and others due to him)
that hold in the Euclidean setting. The next two surveys are concerned with the
geometry of Minkowski spaces (finite-dimensional normed vector spaces). Again,
the results are inspired by analogous results that hold in Euclidean vector spaces.
Then comes a survey on convexity theory, and more precisely on the theory of
mixed volumes for convex bodies in Euclidean space. The goal of the next survey
is to give a set of comparison results and finiteness theorems, in the setting of

A. Papadopoulos ()
Université de Strasbourg and CNRS, Strasbourg, France
e-mail: [email protected]

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 1


A. Papadopoulos (ed.), Surveys in Geometry I,
https://doi.org/10.1007/978-3-030-86695-2_1
2 A. Papadopoulos

metric geometry (more precisely, Gromov hyperbolic spaces), that are analogues
of classical results that hold for Riemannian manifolds. The next two surveys are of
a topological character. A series of techniques are surveyed there, aimed to tackle
classical problems in topology that are related to the 4-dimensional Schoenflies
problem. Translation surfaces with their dynamics, a theory which generalizes in
a very nontrivial manner the classical dynamics of linear flows on the Euclidean
torus, constitute the subject of the next survey. This is followed by a survey which
is concerned with rigidity theorems in Teichmüller spaces, equipped with their two
most important Finsler metrics: the Teichmüller metric and the Thurston metric.
The last survey in this volume is concerned with higher-dimensional complex
geometry, and more especially, with generalizations of classical G-structures and
Cartan geometries.
Each chapter is an illustration of how current research problems and novel
theories are strongly rooted in classical mathematics.
In the rest of this introduction, I review in more detail the content of each of these
chapters.
Chapter 2, by Marek Lassak, is titled Spherical Geometry—A Survey on Width
and Thickness of Convex Bodies. It is a study of the geometry of convex bodies in
d-dimensional spheres. The author develops the spherical analogue of the theory of
width and thickness of convex bodies in Euclidean space. Supporting hemispheres
and lunes play the same role as support hyperplanes and half-spaces in the Euclidean
setting. Among the topics considered in the spherical case, we mention the notions
of diameter, perimeter, circumradius, area, extreme point, reduced body, body of
constant diameter, complete body and body of constant width. The chapter contains
several examples and open questions.
Chapter 3, by Vitor Balestro and Horst Martini, titled Minkowski Geometry—
Some Concepts and Recent Developments, is an introduction to Minkowski geome-
try, that is, the geometry of finite-dimensional normed spaces. The authors study
triangles and the notions of orthocenter, circumcenter, circumradius, Euler line,
Monge point and several others, in a general Minkowski space. They also survey
area and volume, regular polygons, the geometry of circles and that of systems
of circles, Feuerbach circles, equilateral sets and the equilateral dimension of a
normed space, and they review Minkowskian analogues of results on intersections of
circles associated with triangles and circle patterns in the Euclidean plane. They give
several characterizations of Euclidean planes among normed planes. Orthogonality
in Minskowski spaces, also considered in this chapter, is studied in more detail in
the next one. In dimension 2, the notion of anti-norm leads to that of Radon plane
(this is the case where the anti-norm is a multiple of the norm, or, equivalently, when
Birkhoff orthogonality is symmetric).
In the same chapter, the authors study the differential geometry of curves in
Minkowski planes, in particular the notions of Minkowski curvature, radius of
curvature, circular curvature, normal curvature, arclength curvature and curves of
constant width. They survey the differential geometry of surfaces, extending to the
Minkowski setting classical notions such as the Gauss map, principal curvature,
Gaussian curvature, mean curvature, normal curvature, umbilic, Dupin indicatrix,
1 Introduction 3

Dupin metric, minimal surface and girth. They also review Busemann’s work on
isoperimetric problems. Finally, they study billiards in arbitrary convex bodies, a
topic which they call Minkowski billiards. The chapter contains a number of open
problems.
In Chapter 4, authored by Javier Alonso, Horst Martini and Senlin Wu and titled
Orthogonality Types in Normed Linear Spaces, the authors review a large number
of notions of orthogonality in normed vector spaces that generalize the classical
notion of orthogonality in the Euclidean plane or in inner product spaces. Among
the notions of orthogonality that are discussed, we mention Roberts orthogonality,
Birkhoff orthogonality, James or isosceles orthogonality, Pythagorean orthogonal-
ity, Carlsson orthogonality, and there are several others. Symmetry, additivity and
other properties of the various notions of orthogonality are discussed. Naturally,
several characterizations of inner product spaces are obtained. Like in the previous
chapter, a certain number of open problems are discussed.
Chapter 5, by Ivan Izmestiev, is titled Convex Bodies: Mixed Volumes and
Inequalities. The author is motivated by the natural question: How does the volume
of a convex set belonging to a certain class depend on the variables defining it? It
turns out that for what concerns volume of convex bodies, an interesting object to
study is the family of r-neighborhoods of such a body for variable r. A formula
attributed to Jakob Steiner says that the volume of the r-neighborhood of a compact
convex body in Euclidean d-space is a degree-d polynomial in r. In fact, two
original versions of this formula were obtained by Steiner: they concern polytopes
and smooth convex bodies in R3 respectively. The first one involves parameters
such as edge lengths and exterior angles, and the second one involves area and the
principal curvatures of the boundary. Izmestiev, in Chap. 5, reviews these formulae
and their n-dimensional generalizations, leading to the notion of general Steiner
polynomial which establishes a relation between volume and the average volumes
of projections of the convex body to lower-dimensional subspaces. He surveys
several geometrical concepts such as average width and total mean curvature and he
presents n-dimensional generalizations of classical formulae attributed to Cauchy
and Crofton. At the same time, the chapter contains an introduction to some basic
notions in convexity theory, starting with elementary concepts such as support
hyperplane, support function, Minkowski functional, Minkowski sum, the Blaschke
selection theorem, the Hausdorff metric on the space of convex bodies and the
volume function on this space, and continuing with more involved topics such as
mixed volumes and their properties (mixed volume inequalities) and in particular the
Alexandrov–Fenchel and Minkowski inequalities that establish relations between
the volume of the Minkowski sum of two compact convex subsets of Rn and the
volumes of the original two subsets. The author also reviews properties of the
Steiner symmetrization of convex bodies, the Blaschke–Santalo inequality and a
related conjecture by Mahler involving the volume of the polar dual of a convex
body. He mentions the relation with isoperimetric inequalities and the spectrum of
the Laplacian.
Chapter 6, by Gérard Besson and Gilles Courtois, is titled Compactness and
Finiteness Results for Gromov-Hyperbolic Spaces. It is a survey of recent results
4 A. Papadopoulos

obtained by the authors together with Sylvestre Gallot and Andrea Sambusetti in
which they give analogues, in the setting of Gromov-hyperbolic metric spaces, of
classical results due to Bishop and Gromov on the growth of balls in Riemannian
manifolds with bounds on their Ricci curvature or on their entropy. The conclusion
in the results of Bishop and Gromov is formulated as a comparison property: under
the curvature bound condition, the volume of a ball of a certain radius at an arbitrary
point is compared with the volume of a ball of the same radius in a simply-connected
Riemannian manifold of constant sectional curvature. Important consequences of
the main results in the Gromov-hyperbolic setting are obtained in the form of
compactness and finiteness theorems.
In this chapter, the reader is led progressively from elementary notions to deep
results in geometry and topology, illustrating important relations between these
two fields and in particular the restrictions that geometry imposes on topology.
The background material presented includes an introduction to Gromov-hyperbolic
spaces equipped with measures and with isometric group actions, entropy in metric
spaces, CAT(0)-spaces, families of metric spaces endowed with Gromov–Hausdorff
distances, Busemann metric spaces, marked groups, growth of groups, the Margulis
Lemma concerning thin-thick decompositions of manifolds, and systoles.
The next two surveys (Chaps. 7 and 8) are by Valentin Poénaru. Before describing
their content, I would like to say a few words on the notion of geometric simple
connectivity which plays a central role there.
A manifold (of arbitrary dimension, compact or not, possibly with or without
boundary) is said to be geometrically simply connected if it admits a proper Morse
function with no critical points of Morse index 1. Such a notion can also be defined
in the combinatorial category, although it is more complicated to state there; it uses a
handlebody decomposition, and it is a property of the relative positions of 1- and 2-
handles. Roughly speaking, it says that “2-handles cancel 1-handles,” a property that
appears in Smale’s proof of the high-dimensional Poincaré conjecture. Geometric
simple connectivity implies simple connectivity in the usual sense (triviality of
the fundamental group). The converse statement involves delicate questions, and
in all generality it is false; for instance, it is known that it is false in dimension 4,
and it is also false for noncompact manifolds with nonempty boundary. But this
converse is true for instance in the case of compact manifolds of dimension ≥5,
by a result of Smale, a fact which was a crucial step in the latter’s proof of the
high-dimensional Poincaré conjecture. It is also true in dimension 3, by Perelman’s
proof of the Poincaré conjecture in this dimension. Poénaru has developed during
several decades an approach to the 3-dimensional Poincaré conjecture based on 4-
dimensional topological constructions which involve in a crucial way the notion of
geometric simple connectivity for smooth 4-manifolds.
Chapter 7 is titled All 4-Dimensional Smooth Schoenflies Balls Are Geometrically
Simply-Connected—A Fast Survey of the Proof. We recall that a Schoenflies ball is
any one of the two compact bounded smooth manifolds into which an arbitrary
smooth embedding of the sphere S n−1 into S n divides the S n . Poénaru gives in
this chapter a survey of the main steps of the proof of the result stated in the
title. The 4-dimensional smooth Schoenflies problem is in the background. This
problem asks whether any smooth 3-dimensional submanifold of the 4-sphere
1 Introduction 5

which is diffeomorphic to the 3-sphere separates it into two 4-manifolds that are
both diffeomorphic to the 4-ball. The question is motivated by an analogue in
dimension two, where one form of the Schoenflies theorem says that for any simple
closed curve in a 2-sphere, any complementary component of this curve union the
curve itself is homeomorphic to a closed disc. Without further restrictions, the n-
dimensional analogue of this theorem is false.
The same author, in Chap. 8, titled Classical Differential Topology and Non-
commutative Geometry, starts by reviewing some standard constructions in the
theory of smooth 4-manifolds, in particular, spaces that are not geometrically
simply connected. He then surveys some connections between various notions in the
geometry and topology of 4-manifolds. This work is part of the author’s project of
understanding the structure at infinity of noncompact smooth 4-manifolds with non-
empty boundary. He promotes the idea that these spaces lead to non-commutative
spaces in the sense of Connes and that this theory leads to interesting questions in
geometric group theory.
Chapter 9, by Daniel Massart, is titled A Short Introduction to Translation
Surfaces, Veech Surfaces and Teichmüller Dynamics. The author starts with an
exposition of some basic background material, namely, the various ways in which
translation surfaces are defined: planar polygons with sides pairwise identified by
translations, atlases with appropriate transition maps, holomorphic (or Abelian) dif-
ferentials, etc. Half-translation surfaces are associated with quadratic differentials.
Each definition of a translation surface emphasizes a particular point of view on
this theory (combinatorial, geometric or complex-analytic). A translation surface
has an associated flow, defined outside the singular locus. The point of view of
Abelian differential induces another flow parametrized by the circle, namely, turning
the differential by an angle θ ∈ [0, 2π]. The existence of these two dynamical
systems leads to the familiar questions of counting the periodic orbits, studying their
distribution, describing associated invariant measures, etc. After reviewing these
questions, Massart surveys properties that are more specific to translation surfaces,
such as the Veech dichotomy: each direction is either completely periodic (all orbits
are periodic or saddle connections) or uniquely ergodic (all infinite trajectories
are uniformly distributed). An important question is to find classes of surfaces
satisfying this dichotomy. The author then introduces the notions of moduli space of
translation surfaces, of local coordinates given by the relative periods of the Abelian
differentials that define them, and of the stratification of this moduli space by the
type of zeroes of the differentials. He discusses relations with Teichmüller spaces
and with the geodesic flow on moduli spaces. He surveys the basic properties of
Teichmüller discs equipped with their GL+ 2 (R) and Veech group actions and he
reports on McMullen’s classification of GL+ 2 (R)-orbit closures in genus two. The
chapter ends with some notes on what is known in higher genus.
Chapter 10 by Ken’ichi Ohshika is titled Teichmüller Spaces and the Rigidity
of Mapping Class Action. The author reviews several rigidity theorems, first in
the setting of the classical Teichmüller metric, and then in the setting of the more
recently introduced Thurston asymmetric metric. He starts with Royden’s theorem
stating that every isometry of Teichmüller space equipped with its Teichmüller
6 A. Papadopoulos

metric is induced by the action of a unique element of the extended mapping


class group (with few exceptional cases). He reviews at the same time a strongly
related result of Royden, which he calls an infinitesimal rigidity result, stating
that every complex-linear isometry between cotangent spaces of Teichmüller space
at any two points, is a scalar multiple of an action induced by a mapping class.
He presents then an analogue of this infinitesimal rigidity theorem in the setting
of Thurston’s asymmetric metric (recent work by Huang–Ohshika–Papadopoulos).
The work involves a detailed analysis of the combinatorial structure of the unit
sphere in the cotangent space at each point of Teichmüller space. At this occasion,
several features of the Finsler metric geometry of Teichmüller space equipped with
the Teichmüller and the Thurston metrics are highlighted.
Chapter 11, by Indranil Biswas and Sorin Dumitrescu, is titled Holomorphic G-
Structures and Foliated Cartan Geometries on Compact Complex Manifolds. The
subject is holomorphic geometric structures on compact complex manifolds, and
more especially, holomorphic G-structures and holomorphic Cartan geometries and
their generalizations. We recall that a Cartan geometry on a complex manifold
is an infinitesimal structure modeled on a homogeneous space G/H where G is
a complex Lie group and H a closed complex Lie subgroup of G. The precise
definition is given in terms of holomorphic principal bundles. A Cartan geometry
is equipped with a curvature tensor which vanishes when the manifold is modeled
(not only infinitesimally) on G/H . Biswas and Dumitrescu develop the notion of
branched holomorphic Cartan geometry on a complex manifold of any dimension,
and a notion of generalized Cartan geometry, inspired by that of branched pro-
jective structure on Riemann surfaces introduced by R. Mandelbaum. Cartan and
generalized Cartan geometries behave well with respect to holomorphic ramified
maps. The authors then develop foliated versions of branched and generalized
Cartan geometries, providing many examples and presenting classification results
for these geometries. A special emphasis is placed on the GL(2, C) and SL(2, C)
cases. In particular, with any holomorphic SL(2, C) structure on a compact complex
manifold of odd dimension is associated a holomorphic Riemannian metric. In
dimension three the two notions are equivalent. The classification of compact
complex manifolds with holomorphic Riemannian metrics is an open question. In
the last section of this chapter, a certain number of commented open problems
and conjectures on holomorphic G-structures and holomorphic (foliated) Cartan
geometries on compact complex manifolds are presented.
Let me close this introduction by a personal note. Reading the original works
of authors from the past was one of the most important parts of my mathematical
education (I am talking especially of Euler, Lobachevsky and Riemann, but also
of ancient authors like Apollonius, Menelaus and Pappus). But I also think that
reading and writing surveys is a very rewarding activity, and that the mathematical
community is always in need of good surveys of old and recent works.
Another random document with
no related content on Scribd:
his account of his sufferings. Hatred which will refuse both health
and fame, without any loss but that of aged intentions, has begun to
seem a moral falsetto. But who can say, without misgiving, that (on
pagan principles) he is wrong? One person only in Heaven and earth
—Heracles. Every ordinary consideration of public and personal
interest has been put before Philoctetes in vain. But there is another
thought which neither the victim himself nor Odysseus nor
Neoptolemus has strength enough to suggest, or even to remember.
One character alone in Greek story rises above the conception of
personal injury or personal benefit as a motive to action; Heracles is
the great reminder, not so much that wrongs to oneself should be
forgiven, as that life is too short and precious to be wasted on
revenge. Heracles would return a blow with vigour, but a vendetta,
in the light of his career, seems a childish folly. One does not forget
that some pictures of his character (that, for instance, in the
Trachiniæ) belie this conception; but Sophocles here sees fit to
choose a different, and the more usual, view. Suddenly the husk of
selfish spite falls away from the sufferer’s soul. He who has just
promised[354] to use the weapons of Heracles in a private quarrel
and has already attempted[355] so to use them, at last remembers
that they are the rightful instrument of well-doing, and that it was
for such a reason[356] that he received them from the hero at his
passing into glory. Heracles then is introduced as the only person
who can press upon Philoctetes an argument which the cunning of
Odysseus and the candour of Neoptolemus have alike ignored. That
he appears as a deus ex machina is in part accident—he is not
selected by the poet for that reason. But it is a happy accident, for
the glory which envelops him is the visible warrant of his inspiring
behests—anything rather than the sign of overwhelming might
summoned to break a reasonable human resolve. Thus the close of
this play is a real ending, not a breakdown; it is the pagan analogue
of the Quo Vadis legend.
The whole play is an example of intrigue. The episode of the
pseudo-merchant[357] is the most brilliant feat of Sophocles in this
department. It reveals to Philoctetes that he is being pursued by the
Greeks, without arousing his suspicion of Neoptolemus, and so gives
occasion for the transfer of the bow when the sufferer’s fit seizes
him; it conveys a strong reminder of urgency to Neoptolemus; and it
enables Odysseus to learn how his plot progresses. Odysseus merely
by telling of his promise to capture Philoctetes is in a fair way to fulfil
it, as it throws his prey into the arms of Neoptolemus. One apparent
fault of construction is of a type which we have already noted. It is
vital that Philoctetes, before his first consent to leave the island,
should know his friend’s real purpose of taking him to Troy. But why
is he told of it? The merest beginner in duplicity would surely
postpone such a revelation until the victim was at sea. But
Sophocles chooses to tighten his plot up in order to give the
situation in one picture.
Dio Chrysostom in one of his most valuable essays[358] sets up a
comparison between the plays on Philoctetes composed by the three
tragedians. The work of Sophocles is the latest, and two peculiarities
help us to see how far his originality went. Firstly, as a companion to
Odysseus he introduces, not Diomedes as Euripides had done, but a
figure new to the Trojan war, an ingenuous lad whose sympathy
brings out what gentleness remains in the sufferer’s heart. Secondly,
the chorus consists of Greek sailors, not of Lemnian natives as in the
two other playwrights. Sophocles will have no Lemnian visitors
because for him it is a cardinal fact that Philoctetes all these years
has been alone save for a chance ship. Thus we gain for a moment a
glance into the actual thoughts of Sophocles: he has made up his
mind that his Philoctetes must be quite solitary. So essential is this
that he falsifies known facts. Lemnos, he says in the second line of
his play, is “untrodden, uninhabited by men,” whereas, both in the
times supposed and in the poet’s own day, it was a populous place.
This, then, gives an invaluable indication of the extent to which
Sophocles felt himself free to re-model his subject-matter. On the
play itself it throws light. The question is to be studied not from the
point of view of the Greek army, but from that of their potential
helper, soured as he is by a more extreme suffering than Æschylus
and Euripides had imagined.
The picture thus conceived is painted with splendid power.
Romantic desolation makes itself felt in the opening words of
Odysseus, and this sense of the frowning grandeur of nature to
which Philoctetes in his despair appeals[359] is everywhere
associated with the pathos of lonely suffering. “While the mountain
nymph, babbling Echo, appearing afar, makes answer to his bitter
cries.”[360] All that he says, from his first exclamation of joy at
hearing again the Greek language to the noble speech in which he
bids farewell to the bitter home which use has made something like
a friend, is instinct with this mingling of romance and pathos.
Deserted by all men he has yet found companions whom in his
misery he addresses; his hands, his poisoned foot, his eyes, his bow,
and the familiar landscape, vocal with the “bass roar of the sea upon
the headland”.[361] Closer even than these is his eternal unseen
companion, Pain, whom he found at his side on first awakening after
the departure of the Greek host: “When my scrutiny had traversed
all the land, no inhabitant could I find therein save Sorrow; and that,
my son, could be met at every turn”.[362]
It has been suspected that the play contains allusions to
contemporary politics, that the poet is thinking of Alcibiades’ return
from exile. In 410—the year before this play was produced—he had
gained credit from the naval victory of Cyzicus. Some, moreover,
have seen in Odysseus the cynical politician of the day. Other
passages read like criticism of the public “atmosphere” at Athens in
the closing years of the great war. The dramatist is making
deliberate comments on contemporary Athenian politics, but he
assuredly did not choose the whole theme of Philoctetes merely
because of Alcibiades’ restoration.[363]
The Œdipus Coloneus[364] (Οἰδίπους ἐπὶ Κολωνῷ) or Œdipus at
Colonus was according to the customary view produced in 401 b.c.,
three years after the poet’s death, by his namesake and grandson.
The background represents the grove of the Furies at Colonus, a
village near Athens. Œdipus, exiled from Thebes, an aged blind
wanderer, enters led by his daughter Antigone. They obtain the
favour of King Theseus and the citizens, Œdipus promising that after
his death his spirit shall defend Athens. Ismene, his daughter, brings
news that an oracle has said that in the struggle between Thebes
and the Seven led by Polynices, son of Œdipus, that side shall win
which has possession of Œdipus. Both parties are now eager for his
support, but he curses both his sons. Creon, King of Thebes, enters,
and failing to gain aught but reproaches, carries off the two girls,
and is about to seize the father also when he is checked by the
arrival of Theseus, who rescues the maidens. Polynices next comes
to beg his father’s aid, but is sent to his doom with curses. Then a
peal of thunder announces to Œdipus that the moment of his
passing is at hand. He bids farewell to his daughters, and, watched
only by Theseus, descends to the underworld; the place of his burial
is to be known to none save Theseus and his successors. Ismene
and Antigone in vain beg to be shown the spot, and finally Antigone
resolves to seek Thebes that she may reconcile her brothers.
This play is simple in structure, superbly rich in execution. Œdipus
dominates all the scenes, which reveal with piercing intensity his
physical helplessness and the spiritual might which, marked at the
opening, is overwhelming at the close. The poet’s task is not merely
to portray the last hours of a much-tried man, but the novitiate of a
superhuman Power. Œdipus at last reaches peace and a welcome
from the infernal gods—he becomes a δαίμων. The terrific feature is
that even in the flesh he anticipates his dæmonic qualities. In the
interview between him and Polynices, the implacable hatred, the
strength, the prophetic sight of the father, and the hopeless prayers,
the wretchedness, the despair and moral collapse of the doomed
son, are nothing but the presentation in human life of the actual
dæmon’s power as prophesied for future generations. Before the
close we feel that the aged exile’s sufferings, sombre wisdom, and
simple burning emotions have already made him a being of
unearthly powers, sundered from normal humanity; his strange
passing is but the ratification of a spiritual fact already accomplished.
But this weird climax is preceded by an equally wonderful study of
the human Œdipus. The king who appears in the Œdipus Tyrannus
can here still be recognized. Passionate anger still directs much of
his conduct, as friend and foe alike remind[365] him. But even his
faults are mellowed by years and contemplation; his very anger
shows some gleam of a profounder patience. Throughout, the
temper of Œdipus is like that of the heavens above him—gloom cleft
by flashes of insight, indignation, and love. Unlike other aged
sufferers, he does not dwell in the past; unlike the saint and martyr
whom a Christian dramatist might have portrayed, he does not lean
upon a future of glory or happiness. Nor again has he sunk into a
senile acquiescence in the present; he is far from being absorbed by
the loving tendance of his daughters. The centre of his life has
shifted, but not to any period of time—rather to another plane of
being. Still in the flesh, his human emotions as essential as ever, his
life is growing assimilated to the non-human existence of the whole
earth. And so it is that Œdipus meets “death” with cheerfulness; he
is departing to his own place. At the last moment the blind man
leads those who see to the place of his departure. What to them is
dreadful and secret is to him the centre of his longing; the terrific
figures who inhabit his new home are welcome friends—at the
beginning of the play he addresses the Furies themselves as “sweet
daughters of primeval Night”.[366] The whole drama at the end is full
of this sense. In the farewell song of the chorus which commends
the wanderer to the powers of Earth, there is an eerie precision and
picturesqueness in the description of the lower world; the “infernal
moor”[367] and the guardian hound gleam forth for an instant into
strange familiarity.
The other characters are carved, though in lower relief, yet with
richness and vigour. Theseus is the ideal Athenian gentleman,[368]
suddenly called to show pity to a pair of helpless wanderers, then
unexpectedly involved in battle with a neighbour state, and finally
confronted with the most awful mysteries of divine government,
without ever losing his courage or his discretion. Creon and
Polynices, such is the immense understanding of the aged poet,
share too in this nobility of mind. They can face facts; and whether
villains or not, they are men of breeding. The “stranger” who first
accosts Œdipus is a charming embodiment of that local patriotism to
which we shall return in a moment. The two sisters are beautifully
distinguished by the divergent experiences of years. Antigone’s
wandering and hardship have made her the more intense and
passionate; Ismene’s life in Thebes have given her comprehension of
more immediate issues. It is through Antigone, moreover, who
declares that she will seek Thebes and attempt to save her brothers,
that the poet obtains one of his noblest effects. Overwhelming as is
the story of Œdipus, his end does not close all; life goes on to
further mysteries of pain and affection.
On the purely literary side the Œdipus Coloneus is certainly the
greatest and the most typical work of Sophocles. The most
celebrated lyric in Greek is the splendid ode in praise of Colonus
—“our white Colonus; where the nightingale, a constant guest, trills
her clear note in the covert of green glades, dwelling amid the wine-
dark ivy and the god’s inviolate bowers, rich in berries and fruit,
unvisited by sun, unvexed by wind of any storm; where the reveller
Dionysus ever walks the ground, companion of the nymphs that
nursed him,”[369]—and of the whole land with its peculiar glories,
the olive of peace and the steed of war. To this should be added that
address of Œdipus to Theseus concerning the fickleness of all things
earthly which is less the speech of one man than the voice of Life
itself.[370] Noblest of all is the account of Œdipus’ last moments, a
passage which in breathless loveliness, pathos, and religious
profundity is beyond telling flawless and without peer. It is curious
that Sophocles in this work which, more than any other, reveals his
own poetic mastery should have definitely drawn attention to the
power of language. At various crises in the play he speaks of the
“little word”[371] and its potency. Œdipus reflects how his two sons
for lack of a “little word” in his defence have suffered him to be
thrust forth into exile; the nobility of Theseus, the sudden hostility of
Thebes in days to come, the appearance of Polynices, are all matters
of the “little word” which means so much. And in his marvellous
farewell to his daughters, Œdipus speaks of the “one word” which
has made all his sorrows vanish—“love”.
Every master-work of literature has a prophetic quality, and
sending its roots down near to the deepest wells of life is instinct
with unconscious kinships. The Œdipus Coloneus is rich in this final
glory of art. The whole conception of the sufferer, aged and blind
but gifted with spiritual sight, recalls the blind Milton’s sublime
address to the Light which “shines inward”; and the thought adds
charm to Sophocles’ description of the nightingale[372] which

Sings darkling, and in shadiest covert hid


Tunes her nocturnal note.

Again, just as the whole scheme suggests King Lear, so does the
simple vigour of Theseus’ words,[373] when he enters at the terrific
close amid the bellowing of the unnatural tempest:—

πάντα γὰρ θεοῦ


τοιαῦτα χειμάζοντος εἰκάσαι πάρα,

recall the “pelting of this pitiless storm”.[374] So too the divine


summons[375] which comes “many times, and manifold” to Œdipus
brings to mind the call “Samuel! Samuel!” The mystery of Œdipus’
tomb suggests the passing of another august soul: “No man
knoweth of his sepulchre unto this day,”[376] and of Joseph, named
among the faithful, who “when he died, made mention of the
departing of the children of Israel; and gave commandment
concerning his bones”.[377]
This play is deeply religious in the very details of its theme as well
as in its tone. Besides the usual orthodox background there is the
lovely presentation of a minor local worship—the cult of the
Eumenides at Colonus, Sophocles’ native village. The aged poet in
his last years seems to have returned with special affection to the
simple observances which he had learnt as a boy, and which evoke
one of his most characteristic phrases:—[378]

τοιαῦτά σοι ταῦτ’ ἐστίν, ὦ ξέν’, οὐ λόγοις


τιμώμεν’, ἀλλὰ τῇ ξυνουσίᾳ πλέον,

the piercing simplicity of which, as well as the theme, suggests


Wordsworth. The lustral bowls, he is careful to tell us, are “the work
of a skilled craftsman”;[379] he is almost on the point of telling us his
name—one can imagine the boy of eighty years ago gazing on these
cups, and his first sense of beauty in workmanship. Those who
would offer sacrifice will find dwelling on the spot a sacristan to
instruct and aid them.[380] Everything on this ground is both familiar
and hallowed; the mysterious “brazen threshold,” the statues of local
deities and heroes, the hollow pear-tree, and the other local
sanctities are carefully particularized, so that the calm beauty of the
country-side and the terrors of religion are strangely and beautifully
interwoven. As for religion in the broader, profounder sense, we
have as elsewhere a reference of human suffering merely to an
inscrutable, divine purpose.[381] But the dramatist indicates carefully
the contribution of human nature to the fulfilment of the oracles;
Polynices[382] becomes subject to his father’s curse because his own
sense of honour forbids him to relinquish his foredoomed enterprise.
Œdipus himself has attained to a wiser conception of sin[383] than
that which rules him at the close of the Tyrannus. He has acted
wrongly, therefore his suffering is just; but he is morally innocent,
and his past actions afflict him as sorrows, not as crimes.[384]

The fragments of the lost plays are on the whole disappointing; a


large proportion are single rare words quoted by ancient
lexicographers, and most of the rest are short sentences or phrases.
There remain the few longer fragments, to which in recent years
important additions have been made.
It seems that the Triptolemus was his first play, produced in 468
b.c. when the poet was twenty-eight. It would then be one of the

works with which he won his victory[385] over Æschylus, and it bore
marks of the older writer’s influence. The theme is the mission of
Triptolemus, who traversed the earth distributing to men corn, the
gift of Demeter, and founded the mysteries at Eleusis. This topic
gave room for a long geographical passage which recalls those of
the Prometheus. Other early dramas were the Thamyras in which
the dramatist himself took the name-part and played the cithara,
and the Nausicaa or Women Washing wherein Sophocles acted the
part of that princess and gained applause by his skill in a game of
ball. The satyric drama Amphiaraus contained a curious scene
wherein an illiterate man conveyed some name or other word to his
hearers by a dance in which his contortions represented successive
letters. Another satyric play The Mustering of the Greeks (Ἀχαιῶν
Σύλλογος) or the Dinner-Party (Σύνδειπνοι) earned the reprobation
of Cicero[386] apparently for its coarseness, which can still be noted
in the fragments. In The Lovers of Achilles (Ἀχιλλέως ἐρασταί) there
was a passage describing the perplexity of passion, which in its
mannered felicity recalls Swinburne or the Sonnets of Shakespeare:

Love is a sweet perplexity of soul,


Most like the sport of younglings, when the sky
In winter-clearness scatters frost abroad:
They seize a glittering icicle, filled a while
With joy and wonder; but ere long the toy
Melts, and they know not how to grasp it still,
Tho’ loth to cast it from them. So with lovers:
Their yearning passion holds them hour by hour
Poised betwixt boldness and reluctant awe.

The Laocoon, which dealt with a famous episode in the capture of


Troy, supplies a fragment describing Æneas’ escape from the city
with his father upon his shoulders; one or two other passages[387]
besides this recall Vergil’s treatment. Another tragedy from the same
cycle of stories, the Polyxena, is praised by “Longinus”[388] in the
same terms of eulogy as the culmination of the Œdipus Coloneus
itself. The Tereus,[389] to judge from the number of fragments, was
very popular; it dealt with the frightful fable of the Thracian King
Tereus, his wife Procne, and her sister Philomela, all of whom were
at last changed into birds. Aristophanes[390] has an obscure series of
jests about this play and the beak-mask with which Sophocles
“outraged” the Thracian monarch. A solitary relic of the Orithyia tells
how the maiden was carried off by the wind-god Boreas

Unto Earth’s verge, beyond the farthest sea,


Vistas of Heaven, and well-springs of the dark,
To the Sun’s ancient garden.

In 1907 there came to light at Oxyrhynchus in Egypt considerable


fragments[391] of two Sophoclean dramas.
Most of these once formed part of the Ichneutæ (Ἰχνευταί) or
Detectives. Formerly we had only two brief and obscure fragments,
and one word quoted by Athenæus; it was known that the play was
satyric. The theme was quite uncertain; and conjecture[392] is now
shown to have gone quite astray. Sophocles, we find, has
dramatized the myth so admirably treated in the Homeric Hymn to
Hermes. A considerable portion of the work can now be read. The
god Apollo announces that his cattle have been stolen and that he
cannot trace them; he offers a reward to anyone who catches the
thief. Silenus and the chorus of satyrs undertake the quest; they are
the “trackers” from whom the play is named. After a time they spy
the footprints of oxen and exclaim that “some god is leading the
colony”. A noise[393] which they cannot understand is heard behind
the scenes. The numerous tracks now give them trouble; they point
backwards here and there—“an odd confusion must have possessed
the herdsman!” Next the satyrs fall on their faces, to the amazement
of Silenus who likens this “trick of hunting on your stomach” to the
position of “a hedgehog in a bush”. They bid him listen; he
importantly replies that they are not helping “my investigation,” loses
his temper, and roundly reviles their cowardice. They recover
themselves and soon arrive at a cave. Silenus kicks at the door until
the nymph Cyllene comes forth. She protests against their boisterous
behaviour, but is appeased by their apologies. When they ask the
meaning of the strange sound, Cyllene reports the birth of the god
Hermes whom she is tending within, and his amazingly rapid growth.
The noise is produced by the babe from “a vessel filled with pleasure
made from a dead beast”. The “detectives” are still perplexed; what
is this creature? The goddess describes the creature in riddling
language. They make laughably divergent guesses: a cat, a panther,
a lizard, a crab, a big-horned beetle; and at last they are told that
the beast is a tortoise. She describes the delight[394] which the child
draws from his playing. The satyrs inform Cyllene that her nursling is
the thief; she indignantly denies that a son of Zeus can have so
acted, and takes the accusation as a joke. They vigorously repeat
their charge, and begin to quarrel with Cyllene. From this point
onwards practically nothing can be made of the papyrus-scraps,
except that Apollo re-appears, and seems to be giving the
“detectives” their reward.
The papyrus which contained the other play, the Eurypylus,[395] is
in tiny fragments, but some of these, combined with our
independent knowledge of the story, enable us to give an outline of
the plot. Astyoche, mother of Eurypylus, was induced by Priam to
allow her son to help the Trojans against the Greeks. He met
Neoptolemus, son of Achilles, in battle and was slain. A messenger
related the encounter to (it seems) his mother Astyoche. The body
was received by Priam with lamentation as if for a son of his own.
This fragment is much the most striking of the collection.
Sophocles’ position in literary history has already been indicated.
[396] We shall here discuss his mind and his art in general outline. Of
his political opinions little is known. Though his work abounds in
saws of statecraft, these are of quite general application;[397] and it
would be dangerous to declare which side, if any, he took in the
political crises which were so numerous and so grave in fifth-century
Athens; there is perhaps one hint[398] that he did not approve the
ascendancy of Pericles. As for religion, he seems to have accepted
both the orthodox cults of his country and the current beliefs of the
ordinary Athenian with little reserve or none. This brings us at once
to a fact which must not be ignored—the feeling among readers of
our own day, that Sophocles for all his merits is a little too
complacent, too urbane, lacking somehow in profundity and real grip
upon the soul. The answer is that we come to Sophocles pre-
occupied by the religious questionings which fill our own time and
which, moreover, interest both Æschylus and Euripides; but there is
no reason why Sophocles should share our disquiet or that of his
fellow-craftsmen. That which for Æschylus is the foreground of his
work, forms for Sophocles only the background. He is not especially
interested in religion itself, but in humanity. For Æschylus religion is
an affair of the intellect; for Euripides it is an affair of morals; for
Sophocles it belongs to the sphere of emotion. And the two great
instruments with which he constructs his plays are human emotions
and human will. For all the plays which we possess the same genesis
exists: the chief character experiences some mighty appeal to the
emotions—the feeling of self-respect in Ajax and Œdipus at Athens,
of family love in Antigone and Electra, of revengefulness in
Philoctetes, of wifely dignity and affection in Deianira, of pity in
Œdipus at Thebes; and then creates drama by the magnificent
pathetic staunchness wherewith the will, taking its direction from the
emotion so aroused, presses on ruthlessly in its attempt to satisfy
this impulse. Nothing seems so dear to him as a purpose which
flaunts cold reason, the purpose of any others, and indeed every
other emotion save that which has started the action upon its
course. He sets before us a person determined on some striking act,
and subjects him to all conceivable assaults of reason and
preachments on expediency, showing him unbroken throughout. The
onslaughts upon Ajax, Antigone, Philoctetes, Œdipus, are not mere
stage-rhetoric; they are “sound common-sense,” “appeals to one’s
better self”; and no logical denial can be opposed to them. Only one
power in man is able to withstand them—the will, taking its stand
once for all upon some instinct for clear, simple action. If we never
listen to reason we are lost; but if we always listen we are lost
equally. That these heroes of the will so often come to misery or
death matters little; they have saved their souls alive instead of
sinking themselves in a sordid acceptance of a second-hand morality.
Over against these figures, to emphasize their defiant grandeur, the
poet loves to set persons admirable indeed, but more commonplace,
who emerge in the dread hour when the haughty will has brought
ruin, and approve themselves as the pivot of the situation. The hero
is great and strikes the imagination, but it is on the shoulders of
men like Creon in the Œdipus Tyrannus, Odysseus, Hyllus, Theseus
in the Œdipus Coloneus, that the real burden of the world’s work
may be safely cast. None the less he loves Antigone better than he
loves Ismene, Œdipus rather than Theseus. In one place at least, in
his dislike for the “reasonable” spirit of compromise, he suffers
himself a malicious little reductio ad absurdum. When Chrysothemis
finds Electra uttering her resentment at the palace gate she says:—
[399]

Sister! Again? Why standest at the door


Holding this language? Will no span of years
Teach thee at length to grudge thy foolish spleen
Such empty comfort? Yet mine own heart too
Knows how it sorrows for our present state....
I avow
That in thy spirit dwelleth righteousness
Not in my words. Yet, if I would be free
I must in all things bend to those in power.

As for the plots themselves, their main feature is that deliberate


complexity which we have called intrigue and which was made
possible by the poet’s use of a third actor. After the great
achievements of Æschylus it became necessary to add some fresh
kind of interest; this Euripides found in a readjustment of
sympathies, Sophocles in an increase of dramatic thrill. It is an
exciting moment in the Trachiniæ when, just as Deianira is about to
re-enter the palace, the messenger mysteriously draws her apart
and reveals the truth about the captive Iole. The magnificent death-
scene of Ajax is the outcome of the cunning wherewith he has
thrown his friends off the scent. The Electra is full of this method;
the mission of Chrysothemis is turned into a weapon against the
murderess who sent her, and the episode of Orestes’ funeral-urn is a
magnificent piece of dramatic artistry. In the Œdipus Tyrannus the
king brings about his own fatal illumination by sending for the
herdsman. The Philoctetes, above all, is filled with the deliberate
plotting of Odysseus. The marked increase in complexity which
Sophocles’ work thus shows as compared with that of Æschylus is
undoubtedly the chief (perhaps the only) reason for his desertion of
the trilogy form.[400] Side by side with this attention to mechanism is
that curious indifference to the fringes of the plot which we have
had occasion to notice in several places.
Another characteristic of Sophocles is that famous “tragic irony” by
which again he imparts new power to old themes. It turns to
magnificent profit a circumstance which might seem to vitiate
dramatic interest—the fact that the spectator knows the myth and
therefore cannot be taken by surprise. Between an audience which
foresees the event, and the stage-personages who cannot, the
playwright sets up a thrilling interest of suspense. He causes his
characters to discuss the future they expect in language which is
fearfully and exquisitely suitable to the future which actually awaits
them. Ajax, while his madness still afflicts him, stands amid the
slaughtered cattle and proclaims his triumph over the Greek
chieftains, just before he awakes to the truth that by his “triumph”
he has ruined himself. More elaborate is the scene in which Deianira
explains her stratagem of the robe which is to bring back the love of
Heracles. But the Œdipus Tyrannus provides by far the finest
instance. As the king in scene after scene accumulates horror upon
his own unconscious head, the spectator receives, always at the
right moment and in full measure, the impact of increasing disaster.
Yet since his perception is a discovery which he himself has made,
horror is tempered by an intellectual glow, a spiritual exaltation.
In the art of iambic verse Sophocles stands beyond all other
Greeks unrivalled. Beside him Æschylus sounds almost clumsy,
Euripides glib, Aristophanes vulgar. Only Shakespeare has that
complete mastery over every shade of emphasis, every possibility of
grandeur and simple ease alike. The iambic line in Sophocles’ hands
can at will display a haunting romantic loveliness, the profoundest
dignity, the sharpest edges of emotion, or the quiet prose of every
day. Consider the following lines,[401] which begin near the end of
Electra’s long speech of complaint to the chorus:—

ΗΛ. ἐγὼ δ’ Ὀρέστην τῶνδε προσμένουσ’ ἀεὶ


παυστῆρ’ ἐφήξειν ἡ τάλαιν’ ἀπόλλυμαι.
μέλλων γὰρ ἀεὶ δρᾶν τι τὰς οὔσας τέ μου
καὶ τὰς ἀπούσας ἐλπίδας διέφθορεν.
ἐν οὖν τοιούτοις οὔτε σωφρονεῖν, φίλαι,
οὔτ’ εὐσεβεῖν πάρεστιν· ἀλλ’ ἔν τοι κακοῖς
πολλή ’στ’ ἀνάγκη κἀπιτηδεύειν κακά.
ΧΟ. φέρ’ εἰπέ, πότερον ὄντος Αἰγίσθου πέλας
λέγεις τάδ’ ἡμῖν, ἢ βεβῶτος ἐκ δόμων;
ΗΛ. ἦ κάρτα μὴ δόκει μ’ ἄν, εἴπερ ἦν πέλας,
θυραῖον οἰχνεῖν· νῦν δ’ ἀγροῖσι τυγχάνει.
ΧΟ. ἦ κἂν ἐγὼ θαρσοῦσα μᾶλλον ἐς λόγους
τοὺς σοὺς ἱκοίμην, εἴπερ ὧδε ταῦτ’ ἔχει;
ΗΛ. ὡς νῦν ἀπόντος ἱστόρει τί σοι φίλον.

Electra’s speech is solemn poetry. The large number of


spondees[402] (there are three in the last line), the slow elaboration
of the ideas—an elaboration admirably pointed by τοι, which brings
the rhythm almost to a standstill—make a strong contrast with the
following conversation. There the relaxation of the rhythm is
unmistakable; φέρ’ εἰπέ is almost casual in its lightness, and it is at
once followed by a tribrach. The rather odd use of the bare dative
ἀγροῖσι is a delightfully neat tinge of colloquialism, supported by
τυγχάνει. The Philoctetes will repay special study from this point of
view. There is a remarkable tendency to divide[403] lines between
speakers in order to express excitement; this device is elsewhere
very uncommon. From this play we may select one example[404] of
amazing skill in rhythm. Philoctetes is explaining how he contrives to
crawl to and fro in quest of food and the like:—

γαστρὶ μὲν τὰ σύμφορα


τόξον τόδ’ ἐξηύρισκε, τὰς ὑποπτέρους
βάλλον πελείας· πρὸς δὲ τοῦθ’, ὅ μοι βάλοι
νευροσπαδὴς ἄτρακτος, αὐτὸς ἂν τάλας
εἰλυόμην δύστηνος ἐξέλκων πόδα
πρὸς τοῦτ’ ἄν.

The dull repetition of πρὸς τοῦτο and of ἄν; the extremely slow
movement of the penultimate line with its three spondees and the
word-ending at the close of the second foot; above all, the manner
in which the whole dragging sentence leads up to the monosyllable
ἄν, so rare at the end of a sentence, and there stops dead, is a
marvellous suggestion of the lame man’s painful progress and of the
way in which at the end of his endurance he falls prone and spent
upon the object of his endeavour.
Specially striking phrases are not common. Sophocles obtains his
effect not by brilliant strokes of diction, but by the cumulative effect
of a sustained manner. There are such dexterities of course, like
Antigone’s πόθος τοι καὶ κακῶν ἄρ’ ἦν τις,[405] and the cry of Electra
to her brother’s ashes:—[406]

τοίγαρ σὺ δέξαι μ’ ἐς τὸ σὸν τόδε στέγος


τὴν μηδὲν εἰς τὸ μηδέν.

A poet who can, by that infinitesimal change from τὸν μηδέν to τὸ


μηδέν, indicate the very soul of grief, may claim to be one of the
immortal masters of language.
Modern readers find one great fault in this poet—colourlessness,
coldness, an absence of hearty verve; he seems a little too polished
and restrained. The truth is that in Sophocles the Attic spirit finds its
literary culmination. Æschylus lives in the pre-Periclean world;
Euripides is too restless and cosmopolitan to reflect the spirit of one
nation only; Plato and Demosthenes belong to the age of
disillusionment which came after Ægospotami; and Thucydides,
though he shows many Attic qualities, is without limpidity. Anyone,
then, who would understand the Athenian genius as embodied in
letters must read Sophocles. He will find the most useful
commentary in the Parthenon and its friezes, and in the remains of
Greek statuary. One of the most marvellous and precious
experiences in life is to gaze upon works like the so-called Fates in
the British Museum, the Venus of Melos, or the Ludovisi Hera. Many
a casual visitor has glanced for the first time at these works and
known strong disappointment. A mere piece of marble accurately
worked into a female face or figure; majestic to be sure—but is this
all? If he will look again he at last perceives that the stone has put
on, not merely humanity, but immortality. An invisible glow radiates
from it like the odour from a flower. We have never found any name
for it but Beauty. It is indeed the quintessence of loveliness, delicate
as gossamer yet indestructible as granite. So with the tragedies of
Sophocles: it is possible to read the Œdipus Tyrannus in certain
moods and find it mere frigid elegance. But, as with the beauties of
Nature, so with the glories of art, it is the second glance, the
lingering of the eye beyond the careless moment, that surprises
something of the ultimate secret.
For reticence is one of the notes of Athenian art. No writer ever
effected so much with so scanty materials as Sophocles; he carries
the art of masterly omission to its extreme. Shakespeare attempts to
express everything; the mere exuberance of his phraseology is as
wonderful as anything else in his work. But even King Lear or
Hamlet, being written by a man, share the weakness of humanity
and leave the foundation of life undisclosed. Such a disability may
daunt the scientist; it is the salvation of the artist; for the effect of
all art rests on co-operation between the maker and the spectator of
the work. In literature, then, the author knows that he must omit,
and the reader or hearer must supply for himself the contributions of
his own heart and experience. How much then is he to omit? On the
varying answers to that question rest the different forms of literature
and the divergent schools of each form. Sophocles has left more to
his hearer than any other writer in the world. Another note of
Athenian art is simplicity. It is not crudeness, nor naïveté, nor
baldness of style. In a thousand passages of Sophocles, Thucydides,
and Plato, the line between savourless banality and the words they
have written is fine indeed, but that little means a whole world of
art. Many a fine author—Marlowe is a conspicuous example—writes
nobly because he writes violently, or with a conscious effort to soar.
But let him once trip, and he sprawls in bombast or nerveless
garrulity. Simplicity without baldness is the most difficult of all
literary excellences, and is yet achieved everywhere by Sophocles
except when he rises to a different level, of which we shall speak
later.
Such then is the cause of Sophoclean frigidity and lack of colour.
He is led to write so by his Attic frugality and economy of effect, by
his knowledge that his audience can follow him into his rarefied
atmosphere, and by another cause. In our own time men have
looked to art for a “message” from more exciting or more lovely
spheres. We talk of “the literature of escape”; for us art must be an
expanding influence. The Athenian sought in it a concentrating
influence. Each citizen who witnessed the Antigone was a member of
a sovereign assembly; he understood foreign policy at first hand;
war or peace depended upon his voice. Many came to watch the
Ajax who had but a while ago fought at Œnophyta or in Egypt. Such
men did not need “local colour” and exciting technicalities. Their
own lives were full of great events. What they asked of art was
serenity, profundity, to blend their own scattered experiences into
one noble picture of life itself, life made beautiful because so
wonderfully comprehended. This was the function of Sophocles and
his brother-craftsmen.
Beyond the normal lucid beauty of lyrics and dialogue, and beyond
the frequent outpourings of splendid eloquence in long speeches,
there is a still higher level of poetry which should be noted. Now and
again his pages are filled with an unearthly splendour. Reference has
been made before to certain isolated lines which combine utter
simplicity with bewildering charm.[407] But here and there the poet
has given us whole speeches in this divine manner. They are always
a comment on the matter in hand, but they are conceived in the
spirit of one who “contemplates all time and all existence,” who
stands apart from man and sees him in his place amid the workings
of the universe. One of these ethereal utterances is the speech[408]
of Œdipus to Theseus who has expressed his doubt whether Thebes
will ever desert the friendship of Athens; it begins:—

Fair Aigeus’ son, only to gods in heaven


Comes no old age nor death of anything;
All else is turmoiled by our master Time.
The earth’s strength fades and manhood’s glory fades,
Faith dies, and unfaith blossoms like a flower.
And who shall find in the open streets of men
Or secret places of his own heart’s love
One wind blow true for ever?

More personal, but instinct with the same glow of imaginative beauty
is the soliloquy[409] of Ajax when at the point of death. It is in
passages like these that one realizes the value of the restraint which
obtains elsewhere; when the author gives his voice full scope the
effect is heartshaking. Ajax’ appeal to the sun-god to “check his
gold-embossed rein” fills with splendour at a word the heavens
which were lowering with horror. It recalls Marlowe’s lines of the
same type and effect though in different application, which suffuse
the agony of Faust with bitter glory:—

Stand still, you ever-moving spheres of Heaven


That time may cease, and midnight never come!
The greatest achievement of Sophocles was, however, reserved till
the close of his life. The messenger’s speech,[410] narrating the last
moments of Œdipus, is the culmination in Greek of whatever
miracles human language can compass in exciting awe and delight.
The poet has bent all his mastery of tense idiom, of varied and
haunting rhythm, all his instinct for the pathos of life and the
mystery of fate, to produce one mighty uplifting of the hearer into
the region where emotion and intellect are no longer opposed but
mingle into something for which we have no name but “Life”.
CHAPTER V
THE WORKS OF EURIPIDES

Of nearly one hundred dramas composed by Euripides


nineteen[411] have survived. These are now discussed in the
approximate chronological order; the precise date of production is,
however, known in but few cases.
The Alcestis[412] (Ἄλκηστις), acted in 438 b.c., when the poet was
already forty-two years old, is the earliest. It formed the fourth play
of a tetralogy which contained the lost works Women of Crete,
Alcmæon at Psophis, and Telephus. Euripides obtained the second
prize, being vanquished by Sophocles—with what play is not known.
The scene is laid at Pheræ and presents the palace of Admetus, King
of Thessaly. The god Apollo relates how he has induced the Fates to
allow Admetus to escape death on his destined day, if he can find
some one to die in his stead. All refused save his wife, Alcestis,
whose death therefore is to happen this very day. Thanatos (Death)
enters and Apollo in vain asks him to spare the queen; a quarrel
follows, and Apollo departs with threats. The chorus of Pheræan
elders enter and hear, from a servant, of Alcestis’ courageous leave-
takings. Next the queen is borne forth and dies amid the
lamentations of her husband and little son. All save the chorus retire
to prepare for the funeral, when Heracles enters. Admetus comes
forth and insists on making the hero his guest, pretending that it is a
stranger who has died. Heracles is taken to the guest-chamber and
the elders reproach Admetus for his unseasonable hospitality. The
funeral procession is moving forward when Pheres, father of
Admetus, enters to pay his respects to the dead. His son with cold
fury repels him: why did he, an aged man, not consent to die, and
so save Alcestis? A vigorous and coarse altercation follows. When all
have gone the butler enters, complaining of Heracles’ drunken
feasting; the latter soon follows, and is quickly sobered by learning
the truth. He proclaims his intention of rescuing Alcestis from
Thanatos, and hurries away. Admetus returns followed by the
chorus, expressing his utter grief and desolation. Heracles arrives
with a veiled woman, whom he says he has won as a prize at some
athletic contest; he must now depart to fulfil his next “labour”—the
capture of Diomedes’ man-eating steeds—and requests Admetus to
take care of the woman till his return. The king reluctantly consents,
and Heracles unveils her, whereupon Admetus recognizes his wife.
She does not speak, being (as Heracles explains) for three days yet
subject to the infernal deities.[413] The play ends with the joy of
Admetus, a dry remark of Heracles on true hospitality, and a few
lines[414] from the chorus expressing wonder at the mysterious ways
of Heaven.
The Greek introductions to this play contain interesting criticisms:
“the close of the drama is somewhat comic”; “the drama is more or
less satyric, because it ends in joy and pleasure”. These remarks,
coupled with the fact that the Alcestis (as the last play of the
tetralogy) occupied the place of the customary satyric drama, have
caused much discussion. It is enough to say here: first, that the
Alcestis is in no sense a satyric play;[415] second, that it undoubtedly
presents comic features; third, that none the less the work belongs
to the sphere of tragedy. It is sometimes difficult, and often
undesirable, to label dramatic poems too definitely; but we must
certainly avoid the impression that this play is a comedy. It deals
poignantly with the most solemn interests of humanity; the comic
scenes merely show, what is almost as obvious elsewhere, that
Euripides imitates actual life more closely than his two great rivals.
Nothing is gained, however, by ignoring the comic element. The
altercation between Apollo and Thanatos contains much that
surprises us—the wit[416] and the eager, wrangling, bargaining tone
of the dispute. Again the quarrel between Pheres and his son,
admirable in its skilful revelation of character, jars terribly when
enacted over the body of Alcestis. Heracles’ half-tipsy lecture to the
slave shocks us in a demigod about to wrestle with Death himself.
But the whole situation as between Alcestis and Admetus, Admetus
and Heracles, is handled with dignity and extraordinary pathos. The
death scene, especially Admetus’ despairing address to his wife; the
even finer passage when the king returns but shrinks from the cold
aspect of his widowed house; the magnificent and lovely odes,
above all the song which describes the wild beasts of Othrys’ side
sporting to the music of Apollo—these are thoroughly suited to
tragedy.
The plot is apparently[417] quite simple, but one fact should be
mentioned. The rescue of Alcestis is due directly to the drunkenness
of Heracles. He is prevented from learning the facts in an ordinary
way by Admetus; had he behaved normally, he would have left
Pheræ still unenlightened, since Admetus has forbidden[418] his
slave to speak. It is his intoxication alone which goads the butler to
explain.
The character-drawing is skilful, often subtle. Heracles, good-
hearted but somewhat dense, sensual and coarse-fibred, is half-way
between the demigod of the Heracles Furens and the boisterous
glutton of comedy. Capable of splendid impulses, he is yet a
masterpiece of breezy tactlessness, as when with hideous slyness he
suggests to Admetus (in the presence of the restored wife) that the
king may console himself by a new marriage. Pheres and Admetus
are an admirable pair. Both are selfish, Admetus with pathetic
unconsciousness, his father with cynical candour. Pheres is quite
willing to give elaborate honour to the dead woman so long as it
costs little; Admetus—is it true of him that he is ready to utter
splendid heroic speeches so long as the sacrifice is made to save
him? Not so; he feels terribly. But the comparison between father
and son reminds us how easily sentiment can become aged into
etiquette. At present, however, he is a man of generous instincts
—“spoiled”. He needs a salutary upheaval of his home: from afar he
prophesies of Thorvald Helmer in A Doll’s House. Alcestis herself is a
curious study. Innumerable readers have extolled her as one of the
noblest figures in Euripides’ great gallery of heroines; this in spite of
the fact that she is frigid and unimaginative, ungenerous and basely
narrow, in her spiritual and social outlook. One great and noble deed
stands to her credit—she is voluntarily dying to preserve Admetus’
life. Our profound respect Alcestis can certainly claim, but the love
and pity of which so much is said are scarcely due to her. They are
extorted, if at all, by the elaborate exertions of the other characters,
who vie with one another in painting a picture of the tenderness
which has illumined the Pheræan palace like quiet sunshine. But a
dramatist cannot build up a great character by a series of
testimonies from friends. He has undoubtedly portrayed an
interesting personality, as he always does, but to put her beside
creations like Medea and Phædra is merely absurd. From the
beginning of her first intolerable speech[419] we know her for that
frightful figure, the thoroughly good woman with no imagination, no
humour, no insight. One hears much of the failures of Euripides; this
is perhaps a real failure. For we are not to suppose that the rigidity
and coldness of Alcestis are a dexterous stroke of art; it is not his
intention to give a novel, true, and unflattering portrait of a
traditional stage favourite, as he so often delighted to do. Everything
indicates that he wished to make Alcestis sublime and lovable. But
there is a fatal difference between her and the later women.
Euripides has realized her from the outside. He has given us in the
mouths of the other characters warm descriptions of her charm, but
he has not succeeded in drawing a charming woman. She has not
“come alive” in his hands.
The plot of the Alcestis has been studied by the late Dr. Verrall in
an essay[420] of extraordinary skill and interest. He lays special
emphasis on certain peculiar features in the treatment. First,
Heracles is represented as in no way the sublime demigod who
ought to have been depicted, in view of the amazing exploit which
awaits him; the only heroic language put into his mouth is uttered
when he is intoxicated, and the account—if it can be called such—
which he gives later of Alcestis’ deliverance shows a studied lack of
impressiveness. Second, Alcestis is interred with unheard-of speed;
Admetus, seeing her expire, instantly makes ready to convey her
body to the tomb. From these facts in chief and from many details
Dr. Verrall deduces his theory that Alcestis never dies at all. Her
expectation of death (founded on the story about Apollo’s bargain)
and the atmosphere of mourning which hangs over Admetus’ house
and capital on the fatal day, have so wrought upon the queen that
she finally swoons. Later Heracles visits the tomb, finds Alcestis
recovering, and restores her to the king. His annoyance with
Admetus, which leads him to allow his host to “think what he
pleases,” coupled with his own rodomontade at the palace gate,
gave rise to the legend that Heracles fought with Death for a woman
who had actually quitted life. Finally, the quasi-theological prologue,
in which Apollo and Thanatos appear and give warrant to the
orthodox rendering of the story, is a mere figment, revealed as such
to the discreet by its utterly ungodlike tone, and only tacked on to a
quite human drama in order to save the poet from legal indictment
as an enemy of current theology.
This superb essay has met with wide-spread admiration, some
adhesion, much opposition, but no refutation. If we are to judge of
the existing plays as one mass, the examination of outstanding
specimens of rationalism such as the Ion will convince us that the
Alcestis is what Dr. Verrall thought it. But this play does stand apart
from the rest, as do the Rhesus and the Cyclops. However close it
may lie to the Medea in date, it is very early in manner; a capital
instance of this, the character of Alcestis, has already been
mentioned. The best view is, perhaps, that curious features which in
other works might appear so bad as to be evidently intended for
some other than the ostensible purpose, are in this case due to
inexpertness.[421] For example, the extraordinary fact that Alcestis’
rescue is due to nothing but the drunkenness of Heracles, is perhaps
a mere oversight on the poet’s part. Similarly the poorness of the
last scene may be no cunning device, but comparative poverty of
inspiration. It is a tenable view that Euripides intended to write a
quite orthodox treatment of the story, but has only partially
succeeded in reaching the sureness and brilliance of his later
compositions.[422]
The Medea (Μήδεια) was produced in 431 b.c. as the first play of a
tetralogy containing also Philoctetes, Dictys, and the satyric play The
Harvesters (Θερισταί). Euripides obtained only the third prize, and
even Sophocles was second to Euphorion, son of Æschylus. The
scene represents the house of Medea at Corinth. She has come
there with her two young sons, and her husband Jason, whom she
helped to gain the Golden Fleece in Colchis. Jason has become
estranged from Medea, owing to his projected marriage with Glauce,
the daughter of King Creon. At this point the play opens. The aged
nurse of Medea comes forth and, in one of the most celebrated
speeches in Euripides, laments her mistress’ flight from Colchis and
her subsequent troubles; she fears that Medea will seek revenge.
The two boys return from play, attended by their old “pædagogus,”
who informs his fellow-servant that King Creon intends to banish
Medea and her children. The nurse sends them within. The chorus of
Corinthian women enter and inquire after Medea, who comes from
the house in the deepest distress. She speaks with deep feeling
about the sorrows and restraints which society puts upon women,
[423] and after a pathetic description of her own forlorn state, begs
her visitors to aid by silence if she finds any means of revenge. They
have just consented, when Creon appears and orders her to leave
the land on the instant, with her children. When she expostulates,
he explains that he fears her: she is well known as a magician;
moreover, she has uttered threats against himself, his daughter, and
Jason. Medea in vain seeks to escape her reputation for “wisdom”; in
spite of her offer to live quietly in Corinth, Creon repeats his behest.
By urgent pleading, she obtains from him one day’s grace. When the
king has departed, Medea addresses the chorus with fierce triumph:
she now has opportunity for revenge. After considering possible
methods, she decides on poison. But first, what refuge is she to find
when her plot has succeeded? she will wait a little, and if no chance
of safe retirement shows itself, she will attack her foes sword in
hand. The chorus, impressed by her spirit, declare that after all the
centuries during which poets have covered women with infamy, now
at last honour is coming to their sex. They lament the decay of truth
and honour, as shown in Jason’s desertion. Jason enters,
reproaching Medea for her folly in alienating the king, but offering
help to lighten her banishment. Medea falls upon him in a terrible
speech, relating all the benefits she has conferred and the crime she
has committed in his cause. Jason replies that it was the Love-God
which constrained her to help him, nor is he ungrateful. But she has
her reward—a reputation among the Greeks for wisdom. He is
contracting this new marriage to provide for his children; Medea’s
complaints are due to short-sighted jealousy. After a bitter debate, in
which Medea scornfully refuses his aid, he retires. The chorus sing
the dread power of Love, and lament the wreck which it has made of
Medea’s life. A stranger enters—Ægeus, King of Athens, who has
been to Delphi for an oracle which shall remove his childlessness.
Medea begs him to give her shelter in Athens whenever she comes
thither from Corinth; in return for this, she will by her art remove his
childlessness. He consents, and withdraws. Sure of her future,
Medea now triumphantly expounds her plan. She will make a
pretended reconciliation with Jason and beg that her children be
allowed to remain. They are to seek Jason’s bride, bearing presents
in order to win this favour. These gifts will be poisoned; the princess
and all who touch her will perish. Then she will slay her children to
complete the misery of Jason. The chorus in vain protest; she turns
from them and despatches a slave to summon Jason. The choric ode
which follows extols, in lines of amazing loveliness, the glory of
Attica—its atmosphere of wisdom, poetry, and love. But how shall
such a land harbour a murderess? Jason returns, and is greeted by
Medea with a speech of contrition by which he is entirely deceived.
She calls her children forth, and there is a pathetic scene which
affects her, for all her guilty purpose, with genuine emotion. She
puts her pretended plan before Jason, and watches the father depart
with the two boys and their pædagogus carrying the presents. The
ode which follows laments the fatal step that has now been taken.
The pædagogus brings the boys back with news that their sentence
of exile has been remitted, and that the princess has accepted the
gifts. Medea addresses herself to the next task. Now that her plot
against Glauce is in train, the children must die. The famous
soliloquy which follows exhibits the sway alternately exerted over her
by maternal love and the thirst for revenge; after a dreadful struggle
she determines to obey her “passion” and embrace vengeance. The
children are sent within. The next ode is a most painfully real and
intimate revelation of a parent’s anxiety and sorrows. A messenger
hurries up, crying to Medea that she must flee; Creon and his
daughter are both dead. Medea greets his news with cool delight,
braces herself for her last deed, and enters the house. The chorus
utter a desperate prayer to the Sun-god to save his descendants;
but at once the children’s cries are heard. Scarcely have they died
away when Jason furiously enters, followed by henchmen. His chief
thought is to save his children from the vengeance of Creon’s
kinsmen. The chorus at once tell him they are dead, and how. In
frenzy he flings himself upon the door. But he suddenly recoils as the
voice of Medea, clear and contemptuous, descends from the air. She
is seen on high, driving a magic chariot given to her by the Sun-god.
There breaks out a frightful wounding altercation, Jason begging
wildly to be allowed to see and to bury his children’s bodies, Medea
sternly refusing; she will herself bury them beyond the borders of
Corinth. She departs through the air, leaving Jason utterly broken.
The literary history of this play is extremely interesting, though
obscure. First, is it later, or earlier, than the Trachiniæ? One general
idea is common to the two tragedies; but the treatment is utterly
dissimilar, and one may not unreasonably believe that Sophocles has
sought to reprove Euripides, to paint his own conception of a noble
wronged wife, and to show how a woman so placed should demean
herself. Secondly, there is some reason[424] to believe that two
editions of the Medea where for a time in existence. Euripides almost
certainly himself remodelled the work, presumably for a second
“production,” but to what extent it is hard to say. Thirdly, and above
all, there is the question of his originality. The longer Greek
“argument” asserts that he appears to have borrowed the drama
from Neophron and to have introduced alterations. This interesting
problem has been discussed elsewhere.[425] Neophron’s play, if one
is to judge by the style and versification of his brief fragments,
should be regarded as written early in the second half of the fifth
century.
The dramatic structure of the Medea calls for the closest attention.
In Sophocles we have observed how that collision of wills and
emotions, which is always the soul of drama, arises from the
confrontation of two persons. In the present drama that collision
takes place in the bosom of a single person. Sophocles would
probably have given us a Jason whose claim upon our sympathy was
hardly less than that of Medea. Complication, with him, is to be
found in his plots, not in his characters. But here we have a subject
which has since proved so rich a mine of tragic and romantic interest
—the study of a soul divided against itself. Medea’s wrongs, her
passionate resentment, and her plans of revenge do not merely
dominate the play, they are the play from the first line to the close.
Certain real or alleged structural defects should be noted. First, we
observe the incredible part taken by the chorus; they raise not a
finger to stay the designs of Medea upon the king and his daughter;
and we are given no reason to suppose that they are unfriendly to
the royal house. The episode of Ægeus, moreover, is puzzling.
Though quite necessary in view of Medea’s helpless condition and
prepared for by her remarks as to a “tower of refuge,”[426] it is quite
unneeded by one who can command a magic flying chariot.
Moreover, this chariot itself has been often censured, notably by
Aristotle,[427] who regards it as to all intents and purposes a deus ex
machina, and on this ground very properly objects to it.
Dr. Verrall’s[428] theory meets all these difficulties. He supposes
that several of Euripides’ plays were originally written for private
performance. The Medea, so acted, had no obtrusive chorus, and no
miraculous escape of the murderess. To the episode of Ægeus
corresponded a finale in which Medea, by allowing her husband to
bury the bodies of his children, and by instituting the religious rites
referred to in our present text,[429] induced both Jason and the
Corinthians to allow her safe passage to Athens. This view, or a view
essentially resembling it, must be accepted, not so much because of
the absurdity involved (as it appears to us) by the presence of the
chorus, as the utter futility of the Ægeus-scene in the present state
of the text.
The characterization shows Euripides at his best. In the heroine he
gives us the first and possibly the finest of his marvellous studies in
feminine human nature. Alcestis he viewed and described from
without; Medea he has imagined from within. Her passionate love,
which is so easily perverted by brutality into murderous hate, her
pride, will-power, ferocity, and dæmonic energy, are all depicted with
flawless mastery and sympathy. Desperate and cruel as this woman
shows herself, she is no cold-blooded plotter. Creon has heard of her
unguarded threats, and his knowledge wellnigh ruins her project.
Her first words to Jason, “thou utter villain,” followed by a complete
and appalling indictment of his cynicism[430] and ingratitude, are not
calculated to lull suspicion. But however passionate, she owns a
splendid intellect. She faces facts[431] and understands her
weaknesses. When seeking an advantage, she can hold herself
magnificently in hand. The pretended reconciliation with Jason is a
scene of weird thrill for the spectators. Her archness in discussing
his influence over the young princess is almost hideous; and while
she weeps in his arms we remember with sick horror her scornful
words after practising successfully the same arts on the king. Above
all, there is here no petulant railing at “unjust gods,” or “blind fate”.
Her undoing in the past has come from “trust in the words of a man
that is a Greek”;[432] her present murderous rage springs from no
Até but from her own passion (θυμός). The dramatist has set himself
to express human life in terms of humanity.
Jason is a superb study—a compound of brilliant manner, stupidity,
and cynicism. If only his own desires, interests, and comforts are
safe, he is prepared to confer all kinds of benefits. The kindly, breezy
words which he addresses to his little sons must have made
hundreds of excellent fathers in the audience feel for a moment a

You might also like