return

Celebratio Mathematica

Shiing-Shen Chern

The life and mathematics of Shiing-Shen Chern

by Richard S. Palais and Chuu-Lian Terng

Introduction

Many math­em­aticians con­sider Shi­ing-Shen Chern to be the out­stand­ing con­trib­ut­or to re­search in dif­fer­en­tial geo­metry in the second half of the twen­ti­eth cen­tury. Just as geo­metry in the first half-cen­tury bears the in­delible stamp of Élie Cartan, so the seal of Chern ap­pears large on the can­vas of geo­metry that has been painted in the past fifty years. And bey­ond the great re­spect and ad­mir­a­tion that his sci­entif­ic ac­com­plish­ments have brought him, there is also a re­mark­able af­fec­tion and es­teem for Chern on the part of count­less col­leagues, stu­dents, and per­son­al friends. This re­flects an­oth­er as­pect of his ca­reer — the friend­ship, warmth, and con­sid­er­a­tion Chern has al­ways shown to oth­ers throughout a life de­voted as much to help­ing young­er math­em­aticians de­vel­op their full po­ten­tial as to his own re­search.

Our re­count­ing of Chern’s life is in two sec­tions: the first, more bio­graph­ic­al in nature, con­cen­trates on de­tails of his per­son­al and fam­ily his­tory; the second gives a brief re­port on his re­search and its in­flu­ence on the de­vel­op­ment of twen­ti­eth-cen­tury math­em­at­ics.

Our main sources for the pre­par­a­tion of this art­icle were the four volumes of Chern’s se­lec­ted pa­pers [24], [28], [30], [29] pub­lished by Spring­er-Ver­lag, a col­lec­tion of Chern’s Chinese art­icles by Sci­ence Press [27], and many con­ver­sa­tions with Chern him­self.

Early life

Chern was born on Oc­to­ber 28, 1911 in Jia Xin. His fath­er, Bao Zheng Chern, passed the city level Civil Ser­vice ex­am­in­a­tions at the end of the Qing Dyn­asty, and later gradu­ated from Zhe Ji­ang Law School and prac­ticed law. He and Chern’s moth­er, Mei Han, had one oth­er son and two daugh­ters.

Be­cause his grand­moth­er liked to have him at home, Shi­ing-Shen was not sent to ele­ment­ary school, but in­stead learned Chinese at home from his aunt. His fath­er was of­ten away work­ing for the gov­ern­ment, but once when his fath­er was at home he taught Shi­ing-Shen about num­bers, and the four arith­met­ic op­er­a­tions. After his fath­er left, Shi­ing-Shen went on to teach him­self arith­met­ic by work­ing out many ex­er­cises in the three volumes of Bi Shuan Math­em­at­ics. Be­cause of this he eas­ily passed the ex­am­in­a­tion and entered Xiu Zhou School, fifth grade, in 1920.

His fath­er worked for the court in Tianjin and de­cided to move the fam­ily there in 1922. Chern entered Fu Lu­en middle school that year and con­tin­ued to find math­em­at­ics easy and in­ter­est­ing. He worked a large num­ber of ex­er­cises in High­er Al­gebra by Hall and Knight, and in Geo­metry and Tri­go­no­metry by Wentworth and Smith. He also en­joyed read­ing and writ­ing.

1926–30, Nankai University

Chern passed the col­lege en­trance ex­am­in­a­tions in 1926, at age fif­teen, and entered Nankai Uni­versity to study Math­em­at­ics. In the late 1920s there were few math­em­aticians with a PhD de­gree in all of China, but Chern’s teach­er, Lifu Ji­ang, had re­ceived a doc­tor­al de­gree from Har­vard with Ju­li­an Coolidge. Ji­ang had a strong in­flu­ence on Chern’s course of study; he was very ser­i­ous about his teach­ing, giv­ing many ex­er­cises and per­son­ally cor­rect­ing all of them. Nankai provided Chern with an ex­cel­lent edu­ca­tion dur­ing four happy years.

1930–34, Qing Hua graduate school

In the early 1930s, many math­em­aticians with PhD de­grees re­cently earned abroad were re­turn­ing to China and start­ing to train stu­dents. It ap­peared to Chern that this new gen­er­a­tion of teach­ers did not en­cour­age stu­dents to be­come ori­gin­al and strike out on their own, but in­stead set them to work on prob­lems that were fairly routine gen­er­al­iz­a­tions of their own thes­is re­search. Chern real­ized that to at­tain his goal of high qual­ity ad­vanced train­ing in math­em­at­ics he would have to study abroad. Since his fam­ily could not cov­er the ex­pense this would in­volve, he knew that he would re­quire the sup­port of a gov­ern­ment fel­low­ship. He learned that a stu­dent gradu­at­ing from Qing Hua gradu­ate school with suf­fi­ciently dis­tin­guished re­cords could be sent abroad with sup­port for fur­ther study, so, after gradu­at­ing from Nankai in 1930, he took and passed the en­trance ex­am­in­a­tion for Qing Hua gradu­ate school. At that time the four pro­fess­ors of math­em­at­ics at Qing Hua were Qinglai Xiong, Guangy­uan Sun, Wuzhi Yang (C. N. Yang’s fath­er), and Zhi­fan Zheng (Chern’s fath­er-in-law to be), and Chern stud­ied pro­ject­ive dif­fer­en­tial geo­metry with Pro­fess­or Sun.

While at Nankai Chern had taken courses from Ji­ang on the the­ory of curves and sur­faces, us­ing a text­book writ­ten by W. Blasch­ke. Chern had found this deep and fas­cin­at­ing, so when Blasch­ke vis­ited Beijing in 1932, Chern at­ten­ded all of his series of six lec­tures on web geo­metry. In 1934, when Chern gradu­ated from Qing Hua, he was awar­ded a two-year fel­low­ship for study in the United States but, be­cause of his high re­gard for Blasch­ke, he re­ques­ted per­mis­sion from Qing Hua to use the fel­low­ship at the Uni­versity of Ham­burg in­stead. The act­ing chair­man, Pro­fess­or Wuzhi Yang, helped both to ar­range the fel­low­ship for Chern and for his per­mis­sion to use it in Ger­many. This was the year that the Nazis were start­ing to ex­pel Jew­ish pro­fess­ors from the Ger­man uni­versit­ies, but Ham­burg Uni­versity had opened only sev­er­al years be­fore and, per­haps be­cause it was so new, it re­mained re­l­at­ively calm and a good place for a young math­em­atician to study.

1934–36, Hamburg University

Chern ar­rived at Ham­burg Uni­versity in Septem­ber of 1934, and star­ted work­ing un­der Blasch­ke’s dir­ec­tion on ap­plic­a­tions of Cartan’s meth­ods in dif­fer­en­tial geo­metry. He re­ceived the Doc­tor of Sci­ence de­gree in Feb­ru­ary 1936. Be­cause Blasch­ke trav­elled fre­quently, Chern worked much of the time with Blasch­ke’s as­sist­ant, Kähler. Per­haps a the ma­jor in­flu­ence on him while at Ham­burg was Kähler’s sem­in­ar on what is now a known as Cartan–Kähler The­ory. This was then a new the­ory and every­one at the a In­sti­tute at­ten­ded the first meet­ing. By the end of the sem­in­ar only Chern was left, but he felt that he had be­nefited greatly from it. When his two year fel­low­ship ended in the sum­mer of 1936, Chern was offered ap­point­ments at both Qing Hua and Beijing Uni­versity. But he was also offered an­oth­er year of sup­port from The Chinese Cul­ture Found­a­tion and, with the re­com­mend­a­tion of Blasch­ke, he went to Par­is in 1936–37 to work un­der the renowned geo­met­er Élie Cartan.

1936–37, Paris

When Chern ar­rived in Par­is in Septem­ber of 1936, Cartan had so many stu­dents eager to work with him that they lined up to see him dur­ing his of­fice hours. For­tu­nately, after two months Cartan in­vited Chern to see him at home for an hour once every oth­er week dur­ing the re­main­ing ten months he was in Par­is. Chern spent all his ef­forts pre­par­ing for these bi­weekly meet­ings, work­ing very hard and very hap­pily. He learned mov­ing frames, the meth­od of equi­val­ence, more of Cartan–Kähler the­ory, and most a im­port­antly ac­cord­ing to Chern him­self, he learned the math­em­at­ic­al lan­guage and the way of think­ing of Cartan. The three pa­pers he wrote dur­ing this peri­od rep­res­en­ted the fruits of only a small part of the re­search that came out of this as­so­ci­ation with Cartan.

1937–43, Kunming and the Southwest University Consortium

Chern re­ceived an ap­point­ment as Pro­fess­or of Math­em­at­ics at Qing Hua in 1937. But be­fore he could re­turn to China, in­vad­ing Ja­pan­ese forces had touched off the long and tra­gic Sino-Ja­pan­ese war. Qing Hua joined with Pek­ing Uni­versity and Nankai Uni­versity to form a three-uni­versity con­sor­ti­um, first at Chang­sha, and then, be­gin­ning in Janu­ary 1938, at Kun­ming, where it was called the South­w­est As­so­ci­ated Uni­versity. Chern taught at both places. It had an ex­cel­lent fac­ulty, and in par­tic­u­lar Luo­geng Hua was also Pro­fess­or of Math­em­at­ics there. Chern had many ex­cel­lent stu­dents in Kun­ming, some of whom later made sub­stan­tial con­tri­bu­tions to math­em­at­ics and phys­ics. Among these were the math­em­atician H. C. Wang and the No­bel prize-win­ning phys­i­cist C. N. Yang. Be­cause of the war, there was little com­mu­nic­a­tion with the out­side world and the ma­ter­i­al life was mea­ger. But Chern was for­tu­nate enough to have Cartan’s re­cent pa­pers to study, and he im­mersed him­self in these and in his own re­search. The work be­gun dur­ing this dif­fi­cult time would later be­come a ma­jor source of in­spir­a­tion in mod­ern math­em­at­ics.

Chern’s family

In 1937 Chern and Ms. Shih-Ning Cheng be­came en­gaged in Chang­sha, hav­ing been in­tro­duced by Wuzhi Yang. She had re­cently gradu­ated from Dong Wu Uni­versity, where she had stud­ied bio­logy. They were mar­ried in Ju­ly of 1939, and Mrs. Chern went to Shang­hai in 1940 to give birth to their first child, a son Buo Lung. The war sep­ar­ated the fam­ily for six years and they were not re­united un­til 1946. They have a second child, a daugh­ter, Pu (mar­ried to Ching­wu Chu, one of the main con­trib­ut­ors in the de­vel­op­ment of high tem­per­at­ure su­per­con­duct­ors).

The Cherns have had a beau­ti­ful and full mar­riage and fam­ily life. Mrs. Chern has al­ways been at his side and Chern greatly ap­pre­ci­ated her ef­forts to main­tain a se­rene en­vir­on­ment for his re­search. He ex­pressed this in a poem he wrote on her six­tieth birth­day:

Thirty-six years to­geth­er
Through times of hap­pi­ness
And times of worry too.
Time’s pas­sage has no mercy.

We fly the Skies and cross the Oceans
To ful­fill my des­tiny;
Rais­ing the chil­dren fell
En­tirely on your shoulders.

How for­tu­nate I am
To have my works to look back upon,
I feel re­grets you still have chores.

Grow­ing old to­geth­er in El Cer­rito is a bless­ing.
Time passes by,
And we hardly no­tice.

In 1978 Chern wrote in the art­icle “A sum­mary of my sci­entif­ic life and works”:

“I would not con­clude this ac­count without men­tion­ing my wife’s role in my life and work. Through war and peace and through bad and good times we have shared a life for forty years, which is both simple and rich. If there is cred­it for my math­em­at­ic­al works, it will be hers as well as mine.”

1943–45, Institute for Advanced Study at Princeton

By now Chern was re­cog­nized as one of the out­stand­ing math­em­aticians of China, and his work was draw­ing in­ter­na­tion­al at­ten­tion. But he felt un­sat­is­fied with his achieve­ments, and when O. Veblen ob­tained a mem­ber­ship for him at the In­sti­tute for Ad­vanced Study in 1943, he de­cided to go des­pite the great dif­fi­culties of war­time travel. In fact, it re­quired sev­en days for Chern to reach the United States by mil­it­ary air­craft!

This was one of the most mo­ment­ous de­cisions of Chern’s life, for in those next two years in Prin­ceton he was to com­plete some of his most ori­gin­al and in­flu­en­tial work. In par­tic­u­lar, he found an in­trins­ic proof of The Gen­er­al­ized Gauss–Bon­net The­or­em [9], and this in turn lead him to dis­cov­er the fam­ous Chern char­ac­ter­ist­ic classes [10]. In 1945 Chern gave an in­vited hour ad­dress to the Amer­ic­an Math­em­at­ic­al So­ci­ety, sum­mar­iz­ing some of these strik­ing new ad­vances. The writ­ten ver­sion of this talk [11] was an un­usu­ally in­flu­en­tial pa­per, and as Heinz Hopf re­marked in re­view­ing it for Math­em­at­ic­al Re­views it signaled the ar­rival of a new age in glob­al dif­fer­en­tial geo­metry (“Dieser Vor­trag… zeigt, dass wir uns ein­er neuen Epoche in der ‘Dif­fer­en­tial­geo­met­rie im Grossen’ befind­en”).

1946–48, Academia Sinica

Chern re­turned to China in the spring of 1946. The Chinese gov­ern­ment had just de­cided to set up an In­sti­tute of Math­em­at­ics as part of Aca­demia Sin­ica. Lifu Ji­ang was des­ig­nated chair­man of the or­gan­iz­ing com­mit­tee, and he in turn ap­poin­ted Chern as one of the com­mit­tee mem­bers. Ji­ang him­self soon went abroad, and the ac­tu­al work of or­gan­iz­ing the In­sti­tute fell to Chern. At the In­sti­tute, tem­por­ar­ily loc­ated in Shang­hai, Chern em­phas­ized the train­ing of young people. He se­lec­ted the best re­cent un­der­gradu­ates from uni­versit­ies all over China and lec­tured to them twelve hours a week on re­cent ad­vances in to­po­logy. Many of today’s out­stand­ing Chinese math­em­aticians came from this group, in­clud­ing Wen­jun Wu, Shant­ao Liao, Guo Tsai Chen, and C. T. Yang. In 1948 the In­sti­tute moved to Nanjing, and Aca­demia Sin­ica elec­ted eighty-one charter mem­bers, Chern be­ing the young­est of these.

Chern was so in­volved in his re­search and with the train­ing of stu­dents that he paid scant at­ten­tion to the civil war that was en­gulf­ing China. One day however, he re­ceived a tele­gram from J. Robert Op­pen­heimer, then Dir­ect­or of the In­sti­tute for Ad­vanced Study, say­ing “If there is any­thing we can do to fa­cil­it­ate your com­ing to this coun­try please let us know.” Chern went to read the Eng­lish lan­guage news­pa­pers and, real­iz­ing that Nanjing would soon be­come em­broiled in the tur­moil that was rap­idly over­tak­ing the coun­try, he de­cided to move the whole fam­ily to Amer­ica. Shortly be­fore leav­ing China he was also offered a po­s­i­tion at the Tata In­sti­tute in Bom­bay. The Cherns left from Shang­hai on Decem­ber 31, 1948, and spent the Spring Semester at the In­sti­tute in Prin­ceton.

1949–60, Chicago University

Chern quickly real­ized that he would not soon be able to re­turn to China, and so would have to find a per­man­ent po­s­i­tion abroad. At this time, Pro­fess­or Mar­shall Stone of the Uni­versity of Chica­go Math­em­at­ics De­part­ment had em­barked on an ag­gress­ive pro­gram of bring­ing to Chica­go stel­lar re­search fig­ures from all over the world, and in a few years time he had made the Chica­go de­part­ment one of the premi­er cen­ters for math­em­at­ic­al re­search and gradu­ate edu­ca­tion world­wide. Among this group of out­stand­ing schol­ars was Chern’s old friend, An­dré Weil, and in the sum­mer of 1949 Chern too ac­cep­ted a pro­fess­or­ship at the Uni­versity of Chica­go. Dur­ing his el­ev­en years there Chern had ten doc­tor­al stu­dents. He left in 1960 for the Uni­versity of Cali­for­nia at Berke­ley, where he re­mained un­til his re­tire­ment in 1979.

Chern and C. N. Yang

Chern’s pa­per on char­ac­ter­ist­ic classes was pub­lished in 1946 and he gave a one semester course on the the­ory of con­nec­tions in 1949. Yang and Mills pub­lished their pa­per in­tro­du­cing the Yang–Mills the­ory in­to phys­ics in 1954. Chern and Yang were to­geth­er in Chica­go in 1949 and again in Prin­ceton in 1954. They are good friends and of­ten met and dis­cussed their re­spect­ive re­search. Re­mark­ably, neither real­ized un­til many years later that they had been study­ing dif­fer­ent as­pects of the same thing!

1960–79, UC Berkeley

Chern has com­men­ted that two factors con­vinced him to make the move to Berke­ley. One was that the Berke­ley de­part­ment was grow­ing vig­or­ously, giv­ing him the op­por­tun­ity to build a strong group in geo­metry. The oth­er was… the warm weath­er. Dur­ing his years at Berke­ley, Chern dir­ec­ted the thes­is re­search of thirty-one stu­dents. He was also teach­er and ment­or to many of the young postdoc­tor­al math­em­aticians who came to Berke­ley for their first jobs. (This group in­cludes one of the coau­thors of this art­icle; the oth­er was sim­il­arly priv­ileged at Chica­go.) Dur­ing this peri­od the Berke­ley De­part­ment be­came a world-fam­ous cen­ter for re­search in geo­metry and to­po­logy. Al­most all geo­met­ers in the United States, and in much of the rest of the world too, have met Chern and been strongly in­flu­enced by him. He has al­ways been friendly, en­cour­aging, and easy to talk with on a per­son­al level, and since the 1950s his re­search pa­pers, lec­ture notes, and mono­graphs have been the stand­ard source for stu­dents de­sir­ing to learn dif­fer­en­tial geo­metry. When he “re­tired” from Berke­ley in 1979, there was a week long “Chern Sym­posi­um” in his hon­or, at­ten­ded by over three hun­dred geo­met­ers. In real­ity, this was a re­tire­ment in name only; dur­ing the five years that fol­lowed, not only did Chern find time to con­tin­ue oc­ca­sion­al teach­ing as Pro­fess­or Emer­it­us, but he also went “up the hill” to serve as the found­ing dir­ect­or of the Berke­ley Math­em­at­ic­al Sci­ences Re­search In­sti­tute (MSRI).

1981–present, the three institutes

In 1981 Chern, to­geth­er with Calv­in Moore, Is­ad­ore Sing­er, and sev­er­al oth­er San Fran­cisco Bay area math­em­aticians wrote a pro­pos­al to the Na­tion­al Sci­ence Found­a­tion for a math­em­at­ic­al re­search in­sti­tute at Berke­ley. Of the many such pro­pos­als sub­mit­ted, this was one of only two that were even­tu­ally fun­ded by the NSF. Chern be­came the first dir­ect­or of the res­ult­ing Math­em­at­ic­al Sci­ences Re­search In­sti­tute (MSRI), serving in this ca­pa­city un­til 1984. MSRI quickly be­came a highly suc­cess­ful in­sti­tute and many cred­it Chern’s in­flu­ence as a ma­jor factor.

In fact, Chern has been in­stru­ment­al in es­tab­lish­ing three im­port­ant in­sti­tutes of math­em­at­ic­al re­search: The Math­em­at­ic­al In­sti­tute of Aca­demia Sin­ica (1946), The Math­em­at­ic­al Sci­ences Re­search In­sti­tute in Berke­ley, Cali­for­nia (1981), and The Nankai In­sti­tute for Math­em­at­ics in Tianjin, China (1985). It was re­mark­able that Chern did this des­pite a re­luct­ance to get in­volved with de­tails of ad­min­is­tra­tion. In such mat­ters his ad­op­tion of Laozi’s philo­sophy of “Wu Wei” (roughly trans­lated as “Let nature take its course”) seems to have worked ad­mir­ably. Chern has al­ways be­lieved strongly that China could and should be­come a world lead­er in math­em­at­ics. But for this to hap­pen he felt two pre­con­di­tions were re­quired:

  1. The ex­ist­ence with­in the Chinese math­em­at­ic­al com­munity of a group of strong, con­fid­ent, cre­at­ive people, who are ded­ic­ated, un­selfish, and as­pire to go bey­ond their teach­ers, even as they wish their stu­dents to go bey­ond them.

  2. Ample sup­port for ex­cel­lent lib­rary fa­cil­it­ies, re­search space, and com­mu­nic­a­tion with the world-wide math­em­at­ic­al com­munity. (Chern claimed that these re­sources were as es­sen­tial for math­em­at­ics as labor­at­or­ies were for the ex­per­i­ment­al sci­ences).

It was to help in achiev­ing these goals that Chern ac­cep­ted the job of or­gan­iz­ing the math­em­at­ics in­sti­tute of Aca­demia Sin­ica dur­ing 1946 to 1948, and the reas­on why he re­turned to Tianjin to found the Math­em­at­ics In­sti­tute at Nankai Uni­versity after his re­tire­ment in 1984 as dir­ect­or of MSRI.

Dur­ing 1965–76, be­cause of the Cul­tur­al Re­volu­tion, China lost a whole gen­er­a­tion of math­em­aticians, and with them much of the tra­di­tion of math­em­at­ic­al re­search. Chern star­ted vis­it­ing China fre­quently after 1972, to lec­ture, to train Chinese math­em­aticians, and to re­kindle these tra­di­tions. In part be­cause of the strong bonds he had with Nankai Uni­versity, he foun­ded the Nankai Math­em­at­ic­al Re­search In­sti­tute there in 1985. This In­sti­tute has its own hous­ing, and at­tracts many vis­it­ors both from China and abroad. In some ways it is modeled after the In­sti­tute for Ad­vanced Study in Prin­ceton. One of its pur­poses is to have a place where ma­ture math­em­aticians and gradu­ate stu­dents from all of China can spend a peri­od of time in con­tact with each oth­er and with for­eign math­em­aticians, con­cen­trat­ing fully on re­search. An­oth­er is to have an in­spir­ing place in which to work; one that will be an in­cent­ive for the very best young math­em­aticians who get their doc­tor­al de­grees abroad to re­turn home to China.

Honors and awards

Chern was in­vited three times to ad­dress The In­ter­na­tion­al Con­gress of Math­em­aticians. He gave an Hour Ad­dress at the 1950 Con­gress in Cam­bridge, Mas­sachu­setts (the first ICM fol­low­ing the Second World War), spoke again in 1958, at Ed­in­burgh, Scot­land, and was in­vited to give a second Hour Ad­dress at the 1970 ICM in Nice, France. These Con­gresses are held only every fourth year and it is un­usu­al for a math­em­atician to be in­vited twice to give a plen­ary Hour Ad­dress.

Dur­ing his long ca­reer Chern was awar­ded nu­mer­ous hon­or­ary de­grees. He was elec­ted to the US Na­tion­al Academy of Sci­ences in 1961, and re­ceived the Na­tion­al Medal of Sci­ence in 1975 and the Wolf Prize in 1983. The Wolf Prize was in­sti­tuted in 1979 by the Wolf Found­a­tion of Is­rael to hon­or sci­ent­ists who had made out­stand­ing con­tri­bu­tions to their field of re­search. Chern donated the prize money he re­ceived from this award to the Nankai Math­em­at­ic­al In­sti­tute. He is also a for­eign mem­ber of The Roy­al So­ci­ety of Lon­don, Academie Lincei, and the French Academy of Sci­ences. A more com­plete list of the hon­ors he re­ceived can be found in the Cur­riculum Vitae in [28].

An overview of Chern’s research

Chern’s math­em­at­ic­al in­terests have been un­usu­ally wide and far-ran­ging and he has made sig­ni­fic­ant con­tri­bu­tions to many areas of geo­metry, both clas­sic­al and mod­ern. Prin­cip­al among these are:

  • Geo­met­ric struc­tures and their equi­val­ence prob­lems

  • In­teg­ral geo­metry

  • Eu­c­lidean dif­fer­en­tial geo­metry

  • Min­im­al sur­faces and min­im­al sub­man­i­folds

  • Holo­morph­ic maps

  • Webs

  • Ex­ter­i­or Dif­fer­en­tial Sys­tems and Par­tial Dif­fer­en­tial Equa­tions

  • The Gauss–Bon­net The­or­em

  • Char­ac­ter­ist­ic classes

Since it would be im­possible with­in the space at our dis­pos­al to present a de­tailed re­view of Chern’s achieve­ments in so many areas, rather than at­tempt­ing a su­per­fi­cial ac­count of all fa­cets of his re­search, we have elec­ted to con­cen­trate on those areas where the ef­fects of his con­tri­bu­tions have, in our opin­ion, been most pro­found and far-reach­ing. For fur­ther in­form­a­tion con­cern­ing Chern’s sci­entif­ic con­tri­bu­tions the read­er may con­sult the four volume set, Shi­ing-Shen Chern: Se­lec­ted Pa­pers [24], [28], [30], [29]. This in­cludes a Cur­riculum Vitae, a full bib­li­o­graphy of his pub­lished pa­pers, art­icles of com­ment­ary by An­dré Weil and Phil­lip Grif­fiths, and a sci­entif­ic auto­bi­o­graphy in which Chern com­ments briefly on many of his pa­pers.

One fur­ther caveat; the read­er should keep in mind that this is a math­em­at­ic­al bio­graphy, not a math­em­at­ic­al his­tory. As such, it con­cen­trates on giv­ing an ac­count of Chern’s own sci­entif­ic con­tri­bu­tions, men­tion­ing oth­er math­em­aticians only if they were his coau­thors or had some par­tic­u­larly dir­ect and per­son­al ef­fect on Chern’s re­search. Chern was work­ing at the cut­ting edge of math­em­at­ics and there were of course many oc­ca­sions when oth­ers made dis­cov­er­ies closely re­lated to Chern’s and at ap­prox­im­ately the same time. A far longer (and dif­fer­ent) art­icle would have been re­quired if we had even at­temp­ted to ana­lyze such cases. But it is not only for reas­ons of space that we have avoided these is­sues. A full his­tor­ic­al treat­ment cov­er­ing this same ground would be an ex­tremely valu­able un­der­tak­ing, and will no doubt one day be writ­ten. But that will re­quire a ma­jor re­search ef­fort of a kind that neither of the present au­thors has the train­ing or qual­i­fic­a­tions even to at­tempt.

Be­fore turn­ing to a de­scrip­tion of Chern’s re­search, we would like to point out a uni­fy­ing theme that runs through all of it: his ab­so­lute mas­tery of the tech­niques of dif­fer­en­tial forms and his art­ful ap­plic­a­tion of these tech­niques in solv­ing geo­met­ric prob­lems. This was a ma­gic mantle, handed down to him by his great teach­er, Élie Cartan. It per­mit­ted him to ex­plore in depth new math­em­at­ic­al ter­rit­ory where oth­ers could not enter. What makes dif­fer­en­tial forms such an ideal tool for study­ing loc­al and glob­al geo­met­ric prop­er­ties (and for re­lat­ing them to each oth­er) is their two com­ple­ment­ary as­pects. They ad­mit, on the one hand, the loc­al op­er­a­tion of ex­ter­i­or dif­fer­en­ti­ation, and on the oth­er the glob­al op­er­a­tion of in­teg­ra­tion over co­chains, and these are re­lated via Stokes’ The­or­em.

Geometric structures and their equivalence problems

Much of Chern’s early work was con­cerned with vari­ous “equi­val­ence prob­lems”. Ba­sic­ally, the ques­tion is how to de­term­ine ef­fect­ively when two geo­met­ric struc­tures of the same type are “equi­val­ent” un­der an ap­pro­pri­ate group of geo­met­ric trans­form­a­tions. For ex­ample, giv­en two curves in space, when is there a Eu­c­lidean mo­tion that car­ries one onto the oth­er? Sim­il­arly, when are two Rieman­ni­an struc­tures loc­ally iso­met­ric? Clas­sic­ally one tried to as­so­ci­ate with a giv­en type of geo­met­ric struc­ture vari­ous “in­vari­ants”, that is, sim­pler and bet­ter un­der­stood ob­jects that do not change un­der an iso­morph­ism, and then show that cer­tain of these in­vari­ants are a “com­plete set”, in the sense that they de­term­ine the struc­ture up to iso­morph­ism. Ideally one should also be able to spe­cify what val­ues these in­vari­ants can as­sume by giv­ing re­la­tions between them that are both ne­ces­sary and suf­fi­cient for the ex­ist­ence of a struc­ture with a giv­en set of in­vari­ants. The goal is a the­or­em like the el­eg­ant clas­sic paradigm of Eu­c­lidean plane geo­metry, stat­ing that the three side lengths of a tri­angle de­term­ine it up to con­gru­ence, and that three pos­it­ive real num­bers arise as side lengths pre­cisely when each is less than the sum of the oth­er two. For smooth, reg­u­lar space curves the solu­tion to the equi­val­ence prob­lem was known early in the last cen­tury. If to a giv­en space curve \( \sigma(s) \) (para­met­rized by arc length) we as­so­ci­ate its curvature \( \kappa(s) \) and tor­sion \( \tau(s) \), it is easy to show that these two smooth scal­ar func­tions are in­vari­ant un­der the group of Eu­c­lidean mo­tions, and that they uniquely de­term­ine a curve up to an ele­ment of that group. Moreover any smooth real val­ued func­tions \( \kappa \) and \( \tau \) can serve as curvature and tor­sion as long as \( \kappa \) is pos­it­ive. The more com­plex equi­val­ence prob­lem for sur­faces in space had also been solved by the mid 1800s. Here the in­vari­ants turned out to be two smooth quad­rat­ic forms on the sur­face, the first and second fun­da­ment­al forms, of which the first, the met­ric tensor, had to be pos­it­ive def­in­ite and the two had to sat­is­fy the so-called Gauss and Codazzi equa­tions. The so-called “form prob­lem”, that is the loc­al equi­val­ence prob­lem for Rieman­ni­an met­rics, was also solved clas­sic­ally (by Chris­tof­fel and Lipschitz). The solu­tion is still more com­plex and su­per­fi­cially seems to have little in com­mon with the oth­er ex­amples above.

As Chern was start­ing his re­search ca­reer, a ma­jor chal­lenge fa­cing geo­metry was to find what this seem­ingly dis­par­ate class of ex­amples had in com­mon, and thereby dis­cov­er a gen­er­al frame­work for the Equi­val­ence Prob­lem. Cartan saw this clearly, and had already made im­port­ant steps in that dir­ec­tion with his gen­er­al ma­chinery of “mov­ing frames”. His ap­proach was to re­duce a gen­er­al equi­val­ence prob­lem to one of a spe­cial class of equi­val­ence prob­lems for dif­fer­en­tial forms. More pre­cisely, he would as­so­ci­ate to a giv­en type of loc­al geo­met­ric struc­ture in open sets \( U \) of \( \mathbf{R}^n \), an “equi­val­ent” struc­ture, giv­en by spe­cify­ing:

  1. a sub­group \( G \) of \( \mathbf{GL}(n, \mathbf{R}) \),

  2. cer­tain loc­al coframe fields \( \{\theta_i \} \) in open sub­sets \( U \) of \( \mathbf{R}^n \) (i.e., \( n \) lin­early in­de­pend­ent dif­fer­en­tial 1-forms in \( U \)).

The con­di­tion of equi­val­ence for \( \{\theta_i\} \) in \( U \) and \( \{\theta_i^{\ast}\} \) in \( U^{\ast} \) is the ex­ist­ence of a dif­feo­morph­ism \( \varphi \) of \( U \) with \( U^{\ast} \) such that \[ \varphi^{\ast} (\theta_i^{\ast}) = \sum_{i=1}^n a_{ij} \theta_j ,\] where \( (a_{ij} ) \) is a smooth map of \( U \) in­to \( G \). A geo­met­ric struc­ture defined by the choices (1) and (2) is now usu­ally called a “\( G \)-struc­ture”, a name in­tro­duced by Chern in the course of form­al­iz­ing and ex­plic­at­ing Cartan’s ap­proach. For a giv­en geo­met­ric struc­ture one must choose the re­lated \( G \)-struc­ture so that its no­tion of equi­val­ence co­in­cides with that for the ori­gin­ally giv­en geo­met­ric struc­ture, so the in­vari­ants of the \( G \)-struc­ture will also be the same as for the giv­en geo­met­ric struc­ture. In the case of the form prob­lem one takes \( G = \mathbf{O}(n) \), and giv­en a Rieman­ni­an met­ric \( ds^2 \) in \( U \) chooses any \( \theta_i \) such that \[ ds^2 = \sum_{i=1}^n\theta_i^2 \] in \( U \). While not al­ways so ob­vi­ous as in this case (and a real geo­met­ric in­sight is some­times re­quired for their dis­cov­ery) most oth­er nat­ur­al geo­met­ric equi­val­ence prob­lems, in­clud­ing the ones men­tioned above, do ad­mit re­for­mu­la­tion in terms of \( G \)-struc­tures.

But do we gain any­thing be­sides uni­form­ity from such a re­for­mu­la­tion? In fact, we do, for Cartan also de­veloped gen­er­al tech­niques for find­ing com­plete sets of in­vari­ants for \( G \)-struc­tures. Un­for­tu­nately, however, car­ry­ing out this solu­tion of the Equi­val­ence Prob­lem in com­plete gen­er­al­ity de­pends on his power­ful but dif­fi­cult the­ory of Pfaf­fi­an sys­tems in in­vol­u­tion, with its meth­od of pro­long­a­tion, a the­ory not widely known or well un­der­stood even today. In fact, while his pree­m­in­ence as a geo­met­er was clearly re­cog­nized to­wards the end of his ca­reer, many great math­em­aticians con­fessed to find­ing Cartan’s work hard go­ing at best, and few math­em­aticians of his day were able to com­pre­hend fully his more nov­el and in­nov­at­ive ad­vances. For ex­ample, in a re­view of one of his books (Bull. Amer. Math. Soc. vol. 44, p. 601) H. Weyl made this of­ten quoted ad­mis­sion:

“Cartan is un­doubtedly the greatest liv­ing mas­ter in dif­fer­en­tial geo­metry… Nev­er­the­less I must ad­mit that I found the book, like most of Cartan’s pa­pers, hard read­ing…”

Giv­en this well-known dif­fi­culty Cartan had in com­mu­nic­at­ing his more eso­ter­ic ideas, one can eas­ily ima­gine that his im­port­ant in­sights on the Equi­val­ence Prob­lem might have lain bur­ied. For­tu­nately they were spared such a fate.

Re­call that Chern had spent his time at Ham­burg study­ing the Cartan-Kähler a the­ory of Pfaf­fi­an sys­tems with Kähler, and im­me­di­ately after Ham­burg Chern spent a year in Par­is con­tinu­ing his study of these tech­niques with Cartan. Clearly Chern was ideally pre­pared to carry for­ward the at­tack on the Equi­val­ence Prob­lem. In a series of beau­ti­ful pa­pers over the next twenty years not only did he do just that, but he also ex­plained and re­for­mu­lated the the­ory with such clar­ity and geo­met­ric ap­peal that much (though by no means all!) of the the­ory has be­come part of the com­mon world-view of dif­fer­en­tial geo­met­ers, to be found in the stand­ard text­books on geo­metry. Those two dec­ades were also, not co­in­cid­ent­ally, the years that saw the de­vel­op­ment of the the­ory of fiber bundles and of con­nec­tions on prin­cip­al \( G \)-bundles. These the­or­ies were the res­ult of the com­bined re­search ef­forts of many people and had mul­tiple sources of in­spir­a­tion both in to­po­logy and geo­metry. One ma­jor thread in that de­vel­op­ment was Chern’s work on the Equi­val­ence Prob­lem and his re­lated re­search on char­ac­ter­ist­ic classes that grew out of it. In or­der to dis­cuss this im­port­ant work of Chern we must first define some of the con­cepts and nota­tions that he and oth­ers in­tro­duced.

Us­ing cur­rent geo­met­ric ter­min­o­logy, a \( G \)-struc­ture for a smooth n-di­men­sion­al man­i­fold \( M \) is a re­duc­tion of the struc­ture group of its prin­cip­al tan­gent coframe bundle from \( \mathbf{GL}(n, \mathbf{R}) \) to the sub­group \( G \). In par­tic­u­lar, the total space of this re­duc­tion is a prin­cip­al \( G \)-bundle, \( P \), over \( M \) con­sist­ing of the ad­miss­ible coframes \[ \theta = (\theta_1, \dots, \theta_n) ,\] and we can identi­fy the \( G \)-struc­ture with this \( P \). There are \( n \) ca­non­ic­ally defined 1-forms \( \omega_i \) on \( P \); if \[ \Pi : P\rightarrow M \] is the bundle pro­jec­tion, then the value of \( \omega_i \) at \( \theta \) is \( \Pi^{\ast} (\theta_i ) \). The ker­nel of \( D\Pi \) is of course the sub­bundle of the tan­gent bundle \( T\!P \) of \( P \) tan­gent to its fibers, and is usu­ally called the ver­tic­al sub­bundle \( V \). Clearly the ca­non­ic­al forms \( \omega_i \) van­ish on \( V \). The group \( G \) acts on the right on \( P \), act­ing simply trans­it­ively on each fiber, so we can identi­fy the ver­tic­al space \( V_{\theta} \) at any point \( \theta \) with the Lie al­gebra \( L(G) \) of left-in­vari­ant vec­tor fields on \( G \). Now, as Ehresmann first noted, a “con­nec­tion” in Cartan’s sense for the giv­en \( G \)-struc­ture (or as we now say, a \( G \)-con­nec­tion for the prin­cip­al bundle \( P \)) is the same as a “ho­ri­zont­al” sub­bundle \( H \) of \( T\!P \) com­ple­ment­ary to \( V \) and in­vari­ant un­der \( G \). In­stead of \( H \) it is equi­val­ent to con­sider the pro­jec­tion of \( T\!P \) onto \( V \) along \( H \) which, by the above iden­ti­fic­a­tion of \( V_{\theta} \) with \( L(G) \), is an \( L(G) \)-val­ued 1-form \( \omega \) on \( P \), called the “con­nec­tion 1-form”. If we de­note the right ac­tion of \( g \in G \) on \( P \) by \( R_g \), then the in­vari­ance of \( H \) un­der \( G \) trans­lates to the trans­form­a­tion law \[ R^{\ast}_g (\omega) = \operatorname{ad}(g^{-1}) \circ \omega \] for \( \omega \), where \( \operatorname{ad} \) de­notes the ad­joint rep­res­ent­a­tion of \( G \) on \( L(G) \). \( L(G) \)-val­ued forms on \( P \) trans­form­ing in this way are called equivari­ant. Since \( L(G) \) is a sub­al­gebra of the Lie al­gebra \( L(\mathbf{GL}(n, \mathbf{R})) \) of \( n{\times}n \) matrices, we can re­gard \( \omega \) as an \( n{\times}n \) mat­rix-val­ued 1-form on \( P \), or equi­val­ently as a mat­rix \( \omega_{ij} \) of \( n^2 \) real-val­ued 1-forms on \( P \).

If \( \sigma : [0, 1] \rightarrow M \) is a smooth path in \( M \) from \( p \) to \( q \), then the con­nec­tion defines a ca­non­ic­al \( G \)-equivari­ant map \( \pi_{\sigma} \) of the fiber \( P_p \) to the fiber \( P_q \), called par­al­lel trans­la­tion along \( \sigma \); namely \( \pi_{\sigma} (\theta) =\tilde{\sigma}(1) \), where \( \tilde{\sigma} \) is the unique ho­ri­zont­al lift of \( \sigma \) start­ing at \( \theta \). In gen­er­al, par­al­lel trans­la­tion de­pends on the path \( \sigma \), not just on the en­d­points \( p \) and \( q \). If it de­pends only on the ho­mo­topy class of \( \sigma \) with fixed en­d­points, then the con­nec­tion is called “flat”. It is easy to see that this is so if and only if the ho­ri­zont­al sub­bundle \( H \) of \( T\!P \) is in­teg­rable, and us­ing the Frobeni­us in­teg­rabil­ity cri­terion, this trans­lates to \[ d\omega_{ij} =\sum_k \omega_{ik} \wedge \omega_{kj} .\] Thus it is nat­ur­al to define the mat­rix \( \Omega_{ij} \) of so-called curvature forms of the con­nec­tion, (whose van­ish­ing is ne­ces­sary and suf­fi­cient for flat­ness) by \[ d\omega_{ij}= \sum_k \omega_{ik} \wedge \omega_{kj} - \Omega_{ij} \] or, in mat­rix nota­tion, \[ d\omega = \omega \wedge \omega - \Omega .\] Since \( \omega \) is equivari­ant, so is \( \Omega \). Dif­fer­en­ti­at­ing the de­fin­ing equa­tion of the curvature forms gives the Bi­an­chi iden­tity, \[ d\Omega = \Omega \wedge \omega - \omega \wedge \Omega .\] A loc­al cross-sec­tion \( \theta : U \rightarrow P \) is called an “ad­miss­ible loc­al coframe” for the \( G \)-struc­ture, and we can use it to pull back the con­nec­tion forms and curvature forms to forms \( \psi_{ij} \) and \( \Psi_{ij} \) on \( U \). Any oth­er ad­miss­ible coframe field \( \hat{\theta} \) in \( U \) is re­lated to \( \theta \) by a unique “change of gauge”, \( g \) in \( U \) (i.e., a unique map \( g : U \rightarrow G \)) such that \[ \hat{\theta}(x) = R_{g(x)} \theta(x) .\] If we use \( \hat{\theta} \) to also pull back the con­nec­tion and curvature forms to forms \( \hat{\psi} \) and \( \hat{\Psi} \) on \( U \), then, us­ing mat­rix nota­tion, it fol­lows eas­ily from the equivari­ance of \( \omega \) and \( \Omega \) that \[ \hat{\psi} = dg\, g^{-1} + g\psi g^{-1} \quad\text{and}\quad \hat{\Psi} = g\Psi g^{-1} .\]

But where do con­nec­tions fit in­to the Equi­val­ence Prob­lem? While Cartan’s solu­tion to the equi­val­ence prob­lem for \( G \)-struc­tures was com­plic­ated in the gen­er­al case, it be­came much sim­pler for the spe­cial case that \( G \) is the trivi­al sub­group \( \{e\} \). For this reas­on Cartan had de­veloped a meth­od by which one could some­times re­duce a \( G \)-struc­ture on a man­i­fold \( M \) to an \( \{e\} \)-struc­ture on a new man­i­fold ob­tained by “adding vari­ables” cor­res­pond­ing to co­ordin­ates in the group \( G \). Chern re­cog­nized that this new man­i­fold was just the total space \( P \) of the prin­cip­al \( G \)-bundle, and that Cartan’s re­duc­tion meth­od amoun­ted to find­ing an “in­trins­ic \( G \)-con­nec­tion” for \( P \), i.e., one ca­non­ic­ally as­so­ci­ated to the \( G \)-struc­ture. In­deed the ca­non­ic­al 1-forms \( \omega_i \) to­geth­er with a lin­early in­de­pend­ent set of the con­nec­tion forms \( \omega_{ij} \), defined by the in­trins­ic con­nec­tion, give a ca­non­ic­al coframe field for \( P \), which of course is the same as an \( \{e\} \)-struc­ture. Fi­nally, Chern real­ized that in this set­ting one could de­scribe geo­met­ric­ally the in­vari­ants for a \( G \)-struc­ture giv­en by Cartan’s gen­er­al meth­od; in fact they can all be cal­cu­lated from the curvature forms of the in­trins­ic con­nec­tion.

Note that this cov­ers one of the most im­port­ant ex­amples of a \( G \)-struc­ture; namely the case \( G = \mathbf{O}(n) \), cor­res­pond­ing to Rieman­ni­an geo­metry. The in­trins­ic con­nec­tion is of course the “Levi-Civ­ita con­nec­tion”. Moreover, in this case it is also easy to ex­plain how to go on to “solve the form prob­lem”, i.e., to find ex­pli­citly a com­plete set of loc­al in­vari­ants for a Rieman­ni­an met­ric. In fact, they can be taken as the com­pon­ents of the Riemann curvature tensor and its co­v­ari­ant de­riv­at­ives in Rieman­ni­an nor­mal co­ordin­ates. To see this, note first the ob­vi­ous fact that there is a loc­al iso­metry of the Rieman­ni­an man­i­fold \( (M, g) \) with \( (M^{\ast}, g^{\ast}) \) car­ry­ing the or­thonor­mal frame \( e_i \) at \( p \) to \( e^{\ast}_i \) at \( p^{\ast} \) if and only if in some neigh­bor­hood of the ori­gin the com­pon­ents \( g_{ij} (x) \) of the met­ric tensor of \( M \) with re­spect to the Rieman­ni­an nor­mal co­ordin­ates \( x_k \) defined by \( e_i \) are identic­al to the cor­res­pond­ing com­pon­ents \( g^{\ast}_{ij} (x) \) of the met­ric tensor of \( M^{\ast} \) with re­spect to Rieman­ni­an nor­mal co­ordin­ates defined by \( e^{\ast}_i \). The proof is then com­pleted by us­ing the easy, clas­sic­al fact ([e1], Ap­pendix II) that each coef­fi­cient in the Mac­laur­in ex­pan­sion of \( g_{ij} (x) \) can be ex­pressed as a uni­ver­sal poly­no­mi­al in the com­pon­ents of the Riemann tensor and a fi­nite num­ber of its co­v­ari­ant de­riv­at­ives.

Let us de­note by \( N (G) \) the semi­direct product \( G \ltimes \mathbf{R}^n \) of af­fine trans­form­a­tions of \( \mathbf{R}^n \) gen­er­ated by \( G \) and the trans­la­tions. Cor­res­pond­ingly we can “ex­tend” the prin­cip­al \( G \)-bundle \( P \) of lin­ear frames to the as­so­ci­ated prin­cip­al \( N (G) \)-bundle \( N (P ) \) of af­fine frames. Chern noted in [12] that the above tech­nique could be ex­pressed more nat­ur­ally, and could be gen­er­al­ized to a wide class of groups \( G \), if one looked for in­trins­ic \( N (G) \)-con­nec­tions on \( N (P ) \). The curvature of an \( N (G) \)-con­nec­tion on \( N (P ) \) is a two-form \( \Omega \) on \( N (P ) \) with val­ues in the Lie al­gebra \( L(N (G)) \) of \( N (G) \). Now \( L(N (G)) \) splits ca­non­ic­ally as the dir­ect sum of \( L(G) \) and \( L(\mathbf{R}^n) = \mathbf{R}^n \), and \( \Omega \) splits ac­cord­ingly. The \( \mathbf{R}^n \) val­ued part, \( \tau \), of \( \Omega \) is called the tor­sion of the con­nec­tion, and what Chern ex­ploited was the fact that he could in cer­tain cases define “in­trins­ic” \( N (G) \) con­nec­tions by put­ting con­di­tions on \( \tau \). For ex­ample, the Levi-Civ­ita con­nec­tion can be char­ac­ter­ized as the unique \( N (\mathbf{O}(n)) \) con­nec­tion on \( N (P ) \) such that \( \tau = 0 \). In fact, in [12] Chern showed that if the Lie al­gebra \( L(G) \) sat­is­fied a cer­tain simple al­geb­ra­ic con­di­tion (“prop­erty C”) then it was al­ways pos­sible to define an in­trins­ic \( N (G) \) con­nec­tion in this way, and he proved that any com­pact \( G \) sat­is­fies prop­erty C. He also poin­ted out here, from the point of view of Cartan’s the­ory of pseudogroups, why some \( G \)-struc­tures do not ad­mit in­trins­ic con­nec­tions. The pseudogroup of a \( G \)-struc­ture \( \Pi : P \rightarrow M \) is the pseudogroup of loc­al dif­feo­morph­isms of \( M \) whose dif­fer­en­tial pre­serves the sub­bundle \( P \). It is ele­ment­ary that the group of bundle auto­morph­isms of a prin­cip­al \( G \)-bundle that pre­serve a giv­en \( G \)-con­nec­tion is a fi­nite di­men­sion­al Lie group and so a for­tiori the pseudogroup of a \( G \)-struc­ture with a ca­non­ic­ally defined con­nec­tion will be a Lie group. But there are im­port­ant ex­amples of groups \( G \) for which the pseudogroup of a \( G \)-struc­ture is of in­fin­ite di­men­sion. For ex­ample, if \( n = 2m \) and we take \( G = \mathbf{GL}(m, \mathbf{C}) \), then a \( G \)-struc­ture is the same thing as an al­most-com­plex struc­ture, and the group of auto­morph­isms is an in­fin­ite pseudogroup.

Chern solved many con­crete equi­val­ence prob­lems. In [1] and [4] he car­ried this out for the path geo­metry defined by a third or­der or­din­ary dif­fer­en­tial equa­tion. Here the \( G \)-struc­ture is on the con­tact man­i­fold of unit tan­gent vec­tors of \( \mathbf{R}^2 \), and \( G \) is the ten-di­men­sion­al group of circle pre­serving con­tact trans­form­a­tions. In [2], [3] he gen­er­al­ized this to the path geo­metry of sys­tems of \( n \)-th or­der or­din­ary dif­fer­en­tial equa­tions. In [8] he con­siders a gen­er­al­ized pro­ject­ive geo­metry, i.e., the geo­metry of \( (k + 1)(n - k) \)-para­met­er fam­ily of \( k \)-di­men­sion­al sub­man­i­folds in \( \mathbf{R}^n \), and in [6], [7] the geo­metry defined by an \( (n - 1) \)-para­met­er fam­ily of hy­per­sur­faces in \( \mathbf{R}^n \). In [22] (jointly with Moser) and in [23] he con­siders real hy­per­sur­faces in \( \mathbf{C}^n \). This lat­ter re­search played a fun­da­ment­al role in the de­vel­op­ment of the the­ory of CR man­i­folds.

Integral geometry

The group \( G \) of ri­gid mo­tions of \( \mathbf{R}^n \) acts trans­it­ively on vari­ous spaces \( S \) of geo­met­ric ob­jects (e.g., points, lines, af­fine sub­spaces of a fixed di­men­sion, spheres of a fixed ra­di­us) so that these spaces can be re­garded as ho­mo­gen­eous spaces, \( G/H \), and the in­vari­ant meas­ure on \( G \) in­duces an in­vari­ant meas­ure on \( S \). This is the so-called “kin­emat­ic dens­ity”, first in­tro­duced by Poin­caré, and the ba­sic prob­lem of in­teg­ral geo­metry is to ex­press the in­teg­rals of vari­ous geo­met­ric­ally in­ter­est­ing quant­it­ies with re­spect to the kin­emat­ic dens­ity in terms of known in­teg­ral in­vari­ants (see [17]). The simplest ex­ample is Crofton’s for­mula for a plane curve \( C \), \[ \int n (\ell \cap C)\, d\ell= 2L (C) \] where \( L(C) \) is its length, \( n(\ell \cap C) \) is the num­ber of its in­ter­sec­tion points with a line in the plane, and \( d\ell \) is the kin­emat­ic dens­ity on lines. We can in­ter­pret this for­mula as say­ing that the av­er­age num­ber of times that a line meets a curve (i.e., is in­cid­ent with a point on the curve) is equal to twice the length of the curve.

In [5], Chern laid down the found­a­tions for a much gen­er­al­ized in­teg­ral geo­metry. In [e4], An­dré Weil says of this pa­per that:

“… it lif­ted the whole sub­ject at one stroke to a high­er plane than where Blasch­ke’s school had left it, and I was im­pressed by the un­usu­al tal­ent and depth of un­der­stand­ing that shone through it.”

Chern first ex­ten­ded the clas­sic­al no­tion of “in­cid­ence” to a pair of ele­ments from two ho­mo­gen­eous spaces \( G/H \) and \( G/K \) of the same group \( G \). Giv­en \( aH \in G/H \) and \( bK \in G/K \), Chern calls them “in­cid­ent” if \[ aH \cap bK \neq\emptyset .\] This defin­i­tion plays an im­port­ant role in the the­ory of Tits build­ings.

In [13] and [17] Chern ob­tained fun­da­ment­al kin­emat­ic for­mu­las for two sub­man­i­folds in \( \mathbf{R}^n \). The in­teg­ral in­vari­ants in Chern’s for­mula arise nat­ur­ally in Weyl’s for­mula for the volume of a tube \( T_{\rho} \) of ra­di­us \( \rho \) about a \( k \)-di­men­sion­al sub­man­i­fold \( X \) of \( \mathbf{R}^n \). Set­ting \( m = n - k \), Weyl’s for­mula is: \[ V (T_\rho) =\sum_{0\leq i\leq k,\, i \text{ even}} c_i \mu_i (X)\rho^{m+i}. \] Here the \( c_i \) are con­stants de­pend­ing on \( m \) and \( i \), \[ \mu_i(X) = \int_M I_i(\Omega) \] where \( I_i \) is a cer­tain ad­joint in­vari­ant poly­no­mi­al of de­gree \( i/2 \) on the Lie al­gebra of \( \mathbf{O}(n) \), and \( \Omega \) is the curvature form with re­spect to the in­duced met­ric on \( X \). Chern’s for­mula (also dis­covered in­de­pend­ently by Fe­der­er) is: \[ \int \mu_e (M_1 \cap g M_2) \, dg = \sum_{0 \leq i \leq e,\, i \text{ even}} c_i \mu_i (M_1) \mu_{e-i} (M_2), \] where \( M_1 \) and \( M_2 \) are sub­man­i­folds of \( \mathbf{R}^n \) of di­men­sions \( p \) and \( q \) re­spect­ively, \( e \) is even, \( 0 \leq e \) \( \leq p + q - n \), and \( c_i \) are con­stant de­pend on \( n, p, q, e \). Grif­fiths made the fol­low­ing com­ment con­cern­ing this pa­per [e3]:

“Chern’s proof of [this for­mula] ex­hib­its a num­ber of char­ac­ter­ist­ic fea­tures. Of course, one is the use of mov­ing frames…. An­oth­er is that the proof pro­ceeds by dir­ect com­pu­ta­tion rather than by es­tab­lish­ing an elab­or­ate, con­cep­tu­al frame­work; in fact upon closer in­spec­tion there is such a con­cep­tu­al frame­work, as de­scribed in [5], however, the philo­soph­ic­al basis is not isol­ated but is left to the read­er to un­der­stand by see­ing how it op­er­ates in a non­trivi­al prob­lem.”

Euclidean differential geometry

One of the main top­ics in clas­sic­al dif­fer­en­tial geo­metry is the study of loc­al in­vari­ants of sub­man­i­folds in Eu­c­lidean space un­der the group of ri­gid mo­tions, i.e., the equi­val­ence prob­lem for sub­man­i­folds. The solu­tion is clas­sic­al. In fact, the first and second fun­da­ment­al forms, \( I \) and \( \mathit{II} \), and the in­duced con­nec­tion \( \nabla^\nu \) on the nor­mal bundle of a sub­man­i­fold sat­is­fy the Gauss, Codazzi and Ricci equa­tions, and they form a com­plete set of loc­al in­vari­ants for sub­man­i­folds in \( \mathbf{R}^n \). Ex­pli­citly these in­vari­ants are as fol­lows:

  1. \( I \) is the in­duced met­ric on \( M \),

  2. \( \mathit{II} \) is a quad­rat­ic form on \( M \) with val­ues in the nor­mal bundle \( \nu(M ) \) such that, for any unit tan­gent vec­tor \( u \) and unit nor­mal vec­tor \( v \) at \( p \), \[ {\mathit{II}}_v (u) = \langle{\mathit{II}}(u), v\rangle \] is the curvature at \( p \) of the plane curve \( \sigma \) formed by in­ter­sect­ing \( M \) with the plane spanned by \( u \) and \( v \), and

  3. if \( s \) is a smooth nor­mal field then \( \nabla^{\nu} (s) \) is the or­tho­gon­al pro­jec­tion of the dif­fer­en­tial \( ds \) onto the nor­mal bundle \( \nu(M) \).

\( {\mathit{II}}_v = \langle{\mathit{II}},v\rangle \) is called the second fun­da­ment­al form in the dir­ec­tion of \( v \). The self-ad­joint op­er­at­or \( A_v \) cor­res­pond­ing to \( {\mathit{II}}_v \) is called the shape op­er­at­or of \( M \) in the dir­ec­tion \( v \).

Chern’s work in this field in­volved mainly the re­la­tion between the glob­al geo­metry of sub­man­i­folds and these loc­al in­vari­ants. He wrote many im­port­ant pa­pers in the area, but be­cause of space lim­it­a­tions we will con­cen­trate only on the fol­low­ing:

(1)  Min­im­al sur­faces

Since the first vari­ation for the area func­tion­al for sub­man­i­folds of \( \mathbf{R}^n \) is the trace of the second fun­da­ment­al form, a sub­man­i­fold \( M \) of \( \mathbf{R}^n \) is called min­im­al if \[ \operatorname{trace}(\mathit{II}\,) = 0 .\] Let \( \mathbf{Gr}(2, n) \) de­note the Grass­mann man­i­fold of 2-planes in \( \mathbf{R}^n \). The Gauss map \( G \) of a sur­face \( M \) in \( \mathbf{R}^n \) is the map from \( M \) to the Grass­mann man­i­fold \( \mathbf{Gr}(2, n) \) defined by \[ G(x) = \text{the tangent plane to }M\text{ at }x .\] The Grass­mann man­i­fold \( \mathbf{Gr}(2, n) \) can be iden­ti­fied as the hy­per­quad­ric \[ z^2_1 + \dots + z^2_n = 0 \] of \( \mathbf{{CP}}^{n-1} \) (via the map that sends a 2-plane \( V \) of \( \mathbf{R}^n \) to the com­plex line spanned by \( e_1 + ie_2 \), where \( (e_1 , e_2 ) \) is an or­thonor­mal base for \( V \)). Thus \( \mathbf{Gr}(2, n) \) has a com­plex struc­ture. On the oth­er hand, an ori­ented sur­face in \( \mathbf{R}^n \) has a con­form­al and hence com­plex struc­ture through its in­duced Rieman­ni­an met­ric. In [16], Chern proved that an im­mersed sur­face in \( \mathbf{R}^n \) is min­im­al if and only if the Gauss map is an­ti­ho­lo­morph­ic. This the­or­em was proved by Pinl for \( n = 4 \) and is the start­ing point for re­lat­ing min­im­al sur­faces with the value dis­tri­bu­tion the­ory of Nevan­linna, Weyl, and Ahlfors. One of the fun­da­ment­al res­ults of min­im­al sur­face the­ory is the Bern­stein unique­ness the­or­em, which says that a min­im­al graph \( z = f (x, y) \) in \( \mathbf{R}^3 \), defined for all \( (x, y) \in \mathbf{R}^2 \), must be a plane. Note that the im­age of the Gauss map of an en­tire graph lies in a hemi­sphere. Bern­stein’s the­or­em as gen­er­al­ized by Os­ser­man says that if the im­age of the Gauss map of a com­plete min­im­al sur­face of \( \mathbf{R}^3 \) is not dense, then the min­im­al sur­face is a plane. In [16], us­ing a clas­sic­al the­or­em of E. Borel, Chern gen­er­al­ized the Bern­stein–Os­ser­man the­or­em to a dens­ity the­or­em on the im­age of the Gauss map of a com­plete min­im­al sur­face in \( \mathbf{R}^n \), that is not a plane. More re­fined dens­ity the­or­ems were es­tab­lished in [18], a joint pa­per with Os­ser­man.

Mo­tiv­ated by Calabi’s work on min­im­al 2-spheres in \( \mathbf{S}^n \), Chern de­veloped in [20] a gen­er­al form­al­ism for os­cu­lat­ing spaces for sub­man­i­folds. He proved that giv­en a min­im­al sur­face in a space form there is an in­teger \( m \) such that the os­cu­lat­ing spaces of or­der \( m \) are par­al­lel along the sur­face, and gave a com­plete sys­tem of loc­al in­vari­ants, with their re­la­tions. As a con­sequence, he proved an ana­logue of Calabi’s the­or­em: if a min­im­al sphere of con­stant Gaus­si­an curvature \( K \) in a space form of con­stant sec­tion­al curvature \( c \) is not totally geodes­ic, then \[ K = \frac{2c}{m(m + 1)} .\]

(2)  Tight and taut im­mer­sions

We first re­call a the­or­em of Fenchel, proved in 1929: if \( \alpha(s) \) is a simple closed curve in \( \mathbf{R}^3 \), para­met­rized by its arc length, and \( k(s) \) is its curvature func­tion, then \[ \int |k(s)|\,ds \geq 2\pi ,\] and equal­ity holds if and only if \( \alpha \) is a con­vex plane curve. Fary and Mil­nor proved that if \( \alpha \) is knot­ted then this in­teg­ral must be great­er than \( 4\pi \).

In [14] and [15], Chern and Lashof gen­er­al­ized these res­ults to sub­man­i­folds of \( \mathbf{R}^n \). Let \( M^m \) be a com­pact m-di­men­sion­al man­i­fold, \[ f : M \rightarrow \mathbf{R}^n \] an im­mer­sion, \( \nu^1 (M) \) the unit nor­mal sphere bundle of \( M \), and \( dv \) the nat­ur­al volume ele­ment of \( \nu^1 (M ) \). Let \[ N : \nu^1 (M ) \rightarrow \mathbf{S}^{n-1} \] de­note the nor­mal map, i.e., \( N \) maps the unit nor­mal vec­tor \( v \) at \( x \) to the par­al­lel unit vec­tor at the ori­gin. Let \( da \) de­note the volume ele­ment of \( \mathbf{S}^{n-1} \). Then the Lipschitz–Killing curvature \( G \) on \( \nu^1(M ) \) is defined by the equa­tion \[ N^{\ast} (da) = G\, \mathit{dA} ,\] i.e., \( G(v) \) is the ab­so­lute value of the de­term­in­ant of the shape op­er­at­or \( A_v \) of \( M \) along the unit nor­mal dir­ec­tion \( v \). The ab­so­lute total curvature \( \tau (M, f ) \) of the im­mer­sion \( f \) is the nor­mal­ized volume of the im­age of \( N \), \[ \tau (M, f ) =\frac{1}{c_{n-1}}\,\int_{\nu^1(M)} |\operatorname{det}(A_v)|\,dv, \] where \( c_{n-1} \) is the volume of the unit \( (n - 1) \)-sphere. In [14] Chern and Lashof gen­er­al­ized Fenchel’s the­or­em by show­ing that \[ \tau(M, f ) \geq 2 ,\] with equal­ity if and only if \( M \) is a con­vex hy­per­sur­face of an \( (m + 1) \)-di­men­sion­al af­fine sub­space \( V \). In [15] they ob­tained the sharp­er res­ult that \[ \tau (M, f ) \geq \sum\beta_i (M ), \] where \( \beta_i (M ) \) is the \( i \)-th Betti num­ber of \( M \).

An im­mer­sion \( f : M \rightarrow \mathbf{R}^n \) is called tight if \( \tau (M, f ) \) is equal to the in­fim­um, \( \tau (M ) \), of the ab­so­lute total curvature among all the im­mer­sions of M in­to Eu­c­lidean spaces of ar­bit­rary di­men­sions. The study of ab­so­lute total curvature and tight im­mer­sion has be­come an im­port­ant field in sub­man­i­fold geo­metry that has seen many in­ter­est­ing de­vel­op­ments in re­cent years. An im­port­ant step in this de­vel­op­ment is Kuiper’s re­for­mu­la­tion of tight­ness in terms of crit­ic­al point the­ory. He showed that for a giv­en com­pact man­i­fold \( M \), \( \tau (M ) \) is the Morse num­ber \( \gamma \) of \( M \), i.e., the min­im­um num­ber of crit­ic­al points a nonde­gen­er­ate Morse func­tion must have. Moreover, an im­mer­sion of \( M \) is tight if and only if every nonde­gen­er­ate height func­tion has ex­actly \( \tau (M ) = \gamma \) crit­ic­al points. An­oth­er de­vel­op­ment is the concept of taut im­mer­sion in­tro­duced by Ban­choff and CarterWest. An im­mer­sion of \( M \) in­to \( \mathbf{R}^n \) is called taut if every nonde­gen­er­ate Eu­c­lidean dis­tance func­tion from a fixed point in \( \mathbf{R}^n \) to the sub­man­i­fold has ex­actly \( \gamma \) crit­ic­al points. Taut im­plies tight, and moreover a taut im­mer­sion is an em­bed­ding. Taut­ness is in­vari­ant un­der con­form­al trans­form­a­tions, hence us­ing ste­reo­graph­ic pro­jec­tion we may as­sume taut sub­man­i­folds lie in the sphere. Pinkall proved that the tube \( M_\epsilon \) of ra­di­us \( \epsilon \) around a sub­man­i­fold \( M \) in \( \mathbf{R}^n \) is a taut hy­per­sur­face if and only if \( M \) is a taut sub­man­i­fold. In par­tic­u­lar, this gives two facts: one is that the par­al­lel hy­per­sur­face of a taut hy­per­sur­face in \( \mathbf{S}^n \) is again taut, an­oth­er is that to un­der­stand taut sub­man­i­folds it suf­fices to un­der­stand taut hy­per­sur­faces. Since the Lie sphere group (the group of con­tact trans­form­a­tions car­ry­ing spheres to spheres) is gen­er­ated by con­form­al trans­form­a­tions and par­al­lel trans­la­tions, taut­ness is in­vari­ant un­der the Lie sphere group. Note also that the \( \epsilon \)-tube \( M \) of a sub­man­i­fold \( M \) in \( S^n \) is an im­mersed Le­gendre sub­man­i­fold of the con­tact man­i­fold of the unit tan­gent bundle of \( S^n \). Thus taut­ness really should be defined for Le­gendre sub­man­i­folds of the con­tact man­i­fold of the unit tan­gent sphere bundle of \( \mathbf{S}^n \). Chern and Cecil make this concept pre­cise in [26] and lay some of the ba­sic dif­fer­en­tial geo­met­ric ground­work for Lie sphere geo­metry. There are many in­ter­est­ing ex­amples of tight and taut sub­man­i­folds and many in­ter­est­ing the­or­ems con­cern­ing them. But some of the most ba­sic ques­tions are still un­answered; for ex­ample there are no good ne­ces­sary and suf­fi­cient con­di­tions known for a com­pact man­i­folds to be im­mersed in Eu­c­lidean space as a tight or taut sub­man­i­fold, and a com­plete set of loc­al in­vari­ants for Lie sphere geo­metry is yet to be found.

The generalized Gauss–Bonnet theorem

Geo­met­ers tend to make a sharp dis­tinc­tion between “loc­al” and “glob­al” ques­tions, and it is com­mon not only to re­gard glob­al prob­lems as some­how more im­port­ant, but even to con­sider loc­al the­ory “old-fash­ioned” and un­worthy of ser­i­ous ef­fort. Chern however has al­ways main­tained that re­search on these seem­ingly po­lar as­pects of geo­metry must of ne­ces­sity go hand-in-hand; he felt that one could not hope to at­tack the glob­al the­ory of a geo­met­ric struc­ture un­til one un­der­stood its loc­al the­ory (i.e., the equi­val­ence prob­lem), and moreover, once one had dis­covered the loc­al in­vari­ants of a the­ory, one was well on the way to­wards find­ing its glob­al in­vari­ants as well! We shall next ex­plain how Chern came to this con­trary at­ti­tude, for it is an in­ter­est­ing and re­veal­ing story, in­volving the most ex­cit­ing and im­port­ant events of his re­search ca­reer: his dis­cov­ery of an “in­trins­ic proof” of the Gen­er­al­ized Gauss–Bon­net The­or­em and, flow­ing out of that, his solu­tion of the char­ac­ter­ist­ic class prob­lem for com­plex vec­tor bundles by his strik­ing and el­eg­ant con­struc­tion of what are now called “Chern classes” from his fa­vor­ite raw ma­ter­i­al, the curvature forms of a con­nec­tion. The Gauss–Bon­net The­or­em for a closed, two-di­men­sion­al Rieman­ni­an man­i­fold \( M \) was surely one of the high points of clas­sic­al geo­metry, and it was gen­er­ally re­cog­nized that gen­er­al­iz­ing it to high­er di­men­sion­al Rieman­ni­an man­i­folds was a cent­ral prob­lem of glob­al dif­fer­en­tial geo­metry. The the­or­em states that the most ba­sic to­po­lo­gic­al in­vari­ant of \( M \), its Euler char­ac­ter­ist­ic \( \chi (M ) \), can be ex­pressed as \( 1/2\pi \) times the in­teg­ral over M of its most ba­sic geo­met­ric in­vari­ant, the Gaus­si­an curvature func­tion \( K \). Al­though there were many pub­lished proofs of this, Chern re­proved it for him­self by a new meth­od that was very nat­ur­al from a mov­ing frames per­spect­ive. Moreover, un­like the pub­lished proofs, Chern’s had the po­ten­tial to gen­er­al­ize to high­er di­men­sions.

To ex­plain Chern’s meth­od, we start by ap­ply­ing the stand­ard mov­ing frames ap­proach to n-di­men­sion­al ori­ented Rieman­ni­an man­i­folds \( M \), then spe­cial­ize to \( n = 2 \). The ori­ent­a­tion to­geth­er with the Rieman­ni­an struc­ture give an \( \mathbf{SO}(n) \) struc­ture for \( M \). Since the Lie al­gebra \( L(\mathbf{SO}(n)) \) is just the skew-ad­joint \( n{\times}n \) matrices, in the prin­cip­al \( \mathbf{SO}(n) \) bundle \( F (M ) \) of ori­ented or­thonor­mal frames of \( M \), in ad­di­tion to the \( n \) ca­non­ic­al 1-forms \( (\omega_{i}) \), we will have the con­nec­tion 1-forms for the Levi-Civ­ita con­nec­tion, a skew-ad­joint \( n {\times} n \) mat­rix of 1-forms \( \omega_{ij} \), char­ac­ter­ized uniquely by the zero tor­sion con­di­tion, \[ d\omega_i =\sum_j \omega_{ij} \wedge \omega_j .\] The com­pon­ents \( R_{ijkl} \) of the Riemann curvature tensor in the frame \( \omega_i \) are de­term­ined from the curvature forms \( \Omega_{ij} \) by \[ \Omega_{ij} = \frac{1}{2}\sum_{kl} R_{ijkl} \omega_k \wedge \omega_l \] (plus the con­di­tion of be­ing skew-sym­met­ric in \( (i, j) \) and in \( (k, l) \)).

When \( n = 2 \), the Lie al­gebra \( L(\mathbf{SO}(n)) \) is 1-di­men­sion­al; \( \omega_{11} = \omega_{22} = 0 \) and \( \omega_{21} = -\omega_{12} \), so there is only one in­de­pend­ent \( \omega_{ij} \), namely \( \omega_{12} \), and so only one curvature equa­tion, \[ d\omega_{12} = -\Omega_{12} = -R_{1212} \,\omega_1 \wedge \omega_2 .\] Now it is eas­ily seen that \( R_{1212} \) is a con­stant on every fiber \( \Pi^{-1} (x) \), and its value is in fact the Gaus­si­an curvature \( K(x) \). We can identi­fy the area 2-form, \( \mathit{dA} \), on \( M \) with \( -\theta_1 \wedge \theta_2 \), where \( (\theta_1 , \theta_2 ) \) is any ori­ented or­thonor­mal frame, so that \[ \Pi^{\ast}(\mathit{dA}) = -\omega_1 \wedge \omega_2 .\] Thus we can re­write the above curvature equa­tion as a for­mula for the pull-back of the Gauss–Bon­net in­teg­rand, \( K \mathit{dA} \), to \( F (M ) \): \begin{equation} \label{ast} \Pi^\ast (K \mathit{dA}) = d\omega_{12}. \tag{*} \end{equation} In [25] Chern re­marks that, along with zero tor­sion equa­tions, the for­mula \( (\text{*}) \) con­tains

“… all the in­form­a­tion on loc­al Rieman­ni­an geo­metry in two di­men­sions [and] gives glob­al con­sequences as well. A little med­it­a­tion con­vinces one that \( (\text{*}) \) must be the form­al basis of the Gauss–Bon­net for­mula, and this is in­deed the case. It turns out that the proof of the n-di­men­sion­al Gauss–Bon­net for­mula can be based on this idea….”

Chern no­ticed a re­mark­able prop­erty of \( (\text{*}) \). Since the Gauss–Bon­net in­teg­rand is a 2-form on a 2-di­men­sion­al man­i­fold, it is auto­mat­ic­ally closed, and hence its pull-back un­der \( \Pi^{\ast} \) must also be closed. But (ex­cept when \( M \) is a tor­us) \( K \mathit{dA} \) is nev­er ex­act, so we do not ex­pect its pull back to be ex­act. Nev­er­the­less, \( (\text{*}) \) says that it is! This phe­nomen­on of a closed but nonex­act form on the base of a fiber bundle be­com­ing ex­act when pulled up to the total space is called trans­gres­sion. As we shall see, it plays a key rôle in Chern’s proof.

By ele­ment­ary to­po­logy, in the com­ple­ment \( M^{\prime} \) of any point \( p \) of a closed Rieman­ni­an man­i­fold \( M \) one can al­ways define a smooth vec­tor field \( e_1 \) of unit length, and the in­dex of this vec­tor field at \( p \) is \( \chi(M ) \). We will now see how this well-known char­ac­ter­iz­a­tion of the Euler char­ac­ter­ist­ic to­geth­er with the trans­gres­sion for­mula \( (\text{*}) \) leads quickly to Chern’s proof of the Gauss–Bon­net the­or­em for two-di­men­sion­al \( M \). Let \( e_2 \) de­note the unit length vec­tor field in \( M^{\prime} \) mak­ing \( (e_1 , e_2 ) \) an ori­ented frame, and let \( \theta \) de­note the dual coframe field in \( M^{\prime} \). Since \( \Pi \) com­posed with \( \theta \) is the iden­tity map of \( M^{\prime} \), we have \[ d(\theta^{\ast}(\omega_{12} )) = \theta^\ast (d\omega_{12} ) = K \mathit{dA} \] in \( M^{\prime} \), so \[ \int_M K \mathit{dA} = \int_{M^{\prime}} K \mathit{dA} =\int_{M^{\prime}} d(\theta^{\ast} (\omega_{12} )) .\] If we write \( M_\epsilon \) for the com­ple­ment of the open \( \epsilon \)-ball about \( p \), then \[ \int_{M^{\prime}}= \lim_{\epsilon\rightarrow 0}\int_{M_\epsilon} ,\] and by Stokes’ The­or­em, \[ \int_M K \mathit{dA} = \lim_{\epsilon \rightarrow 0} \int_{S_{\epsilon}} \theta^\ast (\omega_{12} ) ,\] where \( S_\epsilon = \partial M_\epsilon \) is the dis­tance sphere of ra­di­us \( \epsilon \) about \( p \). The proof will be com­plete if we can identi­fy the right hand side of the lat­ter equa­tion with \( 2\pi \) times the in­dex of \( \epsilon_1 \) at \( p \).

Choose Rieman­ni­an nor­mal co­ordin­ates in a neigh­bor­hood \( U \) of \( p \) and let \( (\hat{e}_1 , \hat{e}_2 ) \) de­note the loc­al frame field in \( U \) defined by or­thonor­mal­iz­ing the cor­res­pond­ing co­ordin­ate basis vec­tors, and \( \hat{\theta} \) the dual coframe field. If \( \alpha(x) \) de­notes the angle between \( e_1(x) \) and \( \hat{e}_1 (x) \), then we re­call that the stand­ard ex­pres­sion for the in­dex or wind­ing num­ber of \( e_1 \) with re­spect to \( p \) is \[ \frac{1}{2\pi}\int_C d\alpha \] where \( C \) is a small simple closed curve sur­round­ing \( p \); so we will be done if we can show that the right hand side above is equal to \[ \int_{S_\epsilon} d\alpha .\]

Let \( \rho(\alpha) \in \mathbf{SO}(2) \) de­note ro­ta­tion through an angle \( \alpha \). The gauge trans­form­a­tion \( g : U \rightarrow \mathbf{SO}(2) \) from the coframe \( \hat{\theta} \) to the coframe \( \theta \) is just \( g(x) = \rho(\alpha(x)) \), so by the trans­form­a­tion law for pull-backs of con­nec­tion forms noted above, \[ \theta^\ast (\omega_{12} ) = d\alpha + \hat{\theta}^\ast (\omega_{12} ) .\] Thus \( \int_{S_\epsilon} \theta^\ast (\omega_{12} ) \) can be writ­ten as the sum of two terms. The first is the de­sired \( \int_{S_\epsilon} d\alpha \), and the second term, \[ \int_{S_\epsilon} \hat{\theta}^\ast (\omega_{12} ) \] clearly tends to zero with \( \epsilon \) since the in­teg­rand is con­tinu­ous at \( p \), while the length of \( S_\epsilon \) tends to zero.

We now re­turn to the case of a gen­er­al n-di­men­sion­al ori­ented Rieman­ni­an man­i­fold \( M \) and de­vel­op some ma­chinery we will need to ex­plain the re­mark­able res­ults that grew out of this ap­proach to the two-di­men­sion­al Gauss–Bon­net The­or­em.

A ba­sic prob­lem is how to con­struct dif­fer­en­tial forms on \( M \) ca­non­ic­ally from the met­ric. Up in the coframe bundle, \( F (M ) \), there is an easy way to con­struct dif­fer­en­tial forms nat­ur­ally from the met­ric — simply take “poly­no­mi­als” in the curvature forms \( \Omega_{ij} \). Cer­tain forms \( \Lambda \) con­struc­ted this way will “define” a form \( \lambda \) on \( M \) by the re­la­tion \[ \Lambda = \Pi^\ast \lambda ,\] and these are the forms we are after.

To make this pre­cise we con­sider the ring \( \mathcal{R} \) of poly­no­mi­als with real (or com­plex) coef­fi­cients in \( n(n - 1)/2 \) vari­ables \( \{X_{ij}\} \), \( 1 \leq i < j \leq n \). We use mat­rix nota­tion; \( X \) de­notes the \( n {\times} n \) mat­rix \( X_{ij} \) of ele­ments of \( \mathcal{R} \), where \( X_{ji} = -X_{ij} \) for \( i < j \), and \( X_{ii} = 0 \). For \( g \in \mathbf{SO}(n) \), \[ \operatorname{ad}(g)X = gXg^{-1} \] is the mat­rix \[ \sum_{k,l} g_{ik} X_{kl} g_{jl} \] of ele­ments of \( \mathcal{R} \). If for \( g \) in \( \mathbf{SO}(n) \) and \( P \) in \( \mathcal{R} \) we define \( \operatorname{ad}(g)P \) in \( \mathcal{R} \) by \[ \bigl(\operatorname{ad}(g)P \bigr)(X) = P \bigl(\operatorname{ad}(g)X\bigr) ,\] this defines an “ad­joint” ac­tion of \( \mathbf{SO}(n) \) on \( \mathcal{R} \) (by ring auto­morph­isms). The sub­ring of “ad-in­vari­ant” ele­ments of \( \mathcal{R} \) is de­noted by \( \mathcal{R}^{\operatorname{ad}} \). For fu­ture ref­er­ence we note that we can also re­gard \( X \) as rep­res­ent­ing the gen­er­al \( n {\times} n \) skew-sym­met­ric mat­rix, i.e., the gen­er­al ele­ment of the Lie al­gebra \( L(\mathbf{SO}(n)) \), and \( \mathcal{R} \) is just the ring of poly­no­mi­al func­tions on \( L(\mathbf{SO}(n)) \).

The curvature 2-forms \( \Omega_{ij} \), be­ing of even de­gree, com­mute with each oth­er un­der ex­ter­i­or mul­ti­plic­a­tion, so we can sub­sti­tute them in ele­ments \( P \) of \( \mathcal{R} \); if \( P (X) \) is ho­mo­gen­eous of de­gree \( d \) in the \( X_{ij} \), then \( P (\Omega) \) will be a dif­fer­en­tial \( 2d \)-form on \( F (M ) \).

Now let \( \theta \) be a loc­al or­thonor­mal coframe field in an open set \( U \) of \( M \), i.e., a loc­al sec­tion \( \theta : U \rightarrow F (M ) \), and let \( \Psi = \theta^\ast (\Omega ) \) de­note the mat­rix of pulled back curvature forms in \( U \). Since \( \theta^\ast \) is a Grass­mann al­gebra ho­mo­morph­ism, for any \( P \) in \( \mathcal{R} \), \[ \theta^\ast \bigl(P (\Omega)\bigr) = P (\Psi) .\] In par­tic­u­lar for any \( x \) in \( U \) we have \[ \theta^\ast \bigl(P (\Omega )\bigr)_x = P (\Psi_x ) .\] If \( \hat{\Psi} \) is the mat­rix of curvature forms in \( U \) cor­res­pond­ing to some oth­er loc­al coframe field, \( \hat{\theta} \) in \( U \), and \( g : U \rightarrow \mathbf{SO}(n) \) is the change of gauge map­ping \( \theta \) to \( \hat{\theta} \), then as noted \[ \hat{\Psi}_x = \operatorname{ad}(g(x))\Psi ,\] so we find \[ P (\hat{\Psi}_x ) = \bigl(\operatorname{ad}(g(x))P\bigr)(\Psi_x ) .\] Thus in gen­er­al the pulled back form \( P (\Psi) \) de­pends on the choice of \( \theta \) and is only defined loc­ally, in \( U \). However if (and only if ) P is in the sub­ring \( \mathcal{R}^{\operatorname{ad}} \) of \( \operatorname{ad} \)-in­vari­ant poly­no­mi­als, the form \( P (\Psi) \) is a glob­ally well-defined form on \( M \), in­de­pend­ent of the choice of loc­al frame fields \( \theta \) used to pull back the loc­ally defined curvature matrices \( \Psi \). In this case it is clear that \[ \Pi^\ast \bigl(P (\Psi)\bigr) = P (\Omega ) ,\] a re­la­tion that uniquely de­term­ines \( P (\Psi) \).

There are many ways one might at­tempt to gen­er­al­ize the Gauss–Bon­net The­or­em for sur­faces, but per­haps the most ob­vi­ous and nat­ur­al is to as­so­ci­ate with every com­pact, ori­ented, n-di­men­sion­al Rieman­ni­an man­i­fold without bound­ary, \( M \), an \( n \)-form \( \lambda \) on \( M \) that is ca­non­ic­ally defined from the met­ric, and has the prop­erty that \[ \lambda = c_n \chi(M ) ,\] where \( c_n \) is some uni­ver­sal con­stant. If \( n \) is odd then Poin­caré du­al­ity im­plies that \( \chi(M ) = 0 \) when \( M \) is without bound­ary, and since we will only con­sider the closed case here, we will as­sume \( n = 2k \). (On the oth­er hand, for odd-di­men­sion­al man­i­folds with bound­ary, the Gauss–Bon­net The­or­em is in­ter­est­ing and de­cidedly non­trivi­al!). From the above dis­cus­sion it is clear that we should define \( \lambda = P (\Psi) \), where \( P \) is an ad-in­vari­ant poly­no­mi­al, ho­mo­gen­eous of de­gree \( k \) in the \( X_{ij} \). In fact there is an ob­vi­ous can­did­ate for \( P \) — the clas­sic­al Pfaf­fi­an, \( \operatorname{Pf} \), uniquely de­term­ined (up to sign) by the con­di­tion that \[ \operatorname{Pf}(X)^2 = \det(X) \] (cf. [e2], p. 309).

A Gen­er­al­ized Gauss–Bon­net The­or­em had already been proved in two pa­pers, one by Al­lendo­er­fer and the oth­er by Fenchel. Both proofs were “ex­trins­ic” — they as­sumed \( M \) could be iso­met­ric­ally em­bed­ded in some Eu­c­lidean space. (A pa­per of Al­lendo­er­fer and Weil im­plied that the ex­ist­ence of loc­al iso­met­ric em­bed­dings was enough, thereby set­tling the case of ana­lyt­ic met­rics). These earli­er proofs wrote the Gen­er­al­ized Gauss–Bon­net in­teg­rand as the volume ele­ment times a scal­ar that was a com­plic­ated poly­no­mi­al in the com­pon­ents of the Riemann tensor. In [9] Chern for the first time wrote the in­teg­rand as the Pfaf­fi­an of the curvature forms and then provided a simple and el­eg­ant in­trins­ic proof of the the­or­em along the lines of the above proof for sur­faces.

Let \( \mathbf{S}(M ) \) de­note the bundle of unit vec­tors of the tan­gent bundle to \( M \), and \( \pi : \mathbf{S}(M ) \rightarrow M \) the nat­ur­al pro­jec­tion. Giv­en a coframe \( \theta \) in \( F (M ) \) let \( e_1 (\theta) \) de­note the first ele­ment of the frame dual to \( \theta \). Then \( e_1 : F (M ) \rightarrow \mathbf{S}(M ) \) is a fiber bundle and clearly \( \Pi : F (M ) \rightarrow M \) factors as \[ \Pi = \pi \circ e_1 .\] Let \( \lambda \) be the \( n \)-form \( \operatorname{Pf}(\Psi) \) on \( M \), and \( \Lambda = p^\ast (\lambda) \) its pull-back to \( \mathbf{S}(M ) \). In [9] Chern first proves a trans­gres­sion lemma for \( \Lambda \), i.e., he ex­pli­citly finds an \( (n - 1) \)-form \( \Theta \) on \( \mathbf{S}(M ) \) sat­is­fy­ing \( d\Theta = \Lambda \). As in two di­men­sions let \( M^{\prime} \) be the com­ple­ment of some point \( p \) in \( M \) and con­struct a smooth cross-sec­tion \( \xi \) of \( \mathbf{S}(M ) \) over \( M \). Then \( \pi\circ \xi \) is the iden­tity map of \( M^{\prime} \), so just as in the two di­men­sion­al ar­gu­ment we find \[ d(\xi^\ast (\theta)) = \lambda ,\] and \[ \int_M = \lim_{\epsilon \rightarrow\theta}\int_{S_\epsilon} \xi^\ast (\Theta) .\] Fi­nally, the con­struc­tion of \( \Theta \) is so ex­pli­cit that Chern is able to eval­u­ate the right hand side by an ar­gu­ment sim­il­ar to the one in the sur­face case, and he finds that it is in­deed a uni­ver­sal con­stant times the Euler char­ac­ter­ist­ic of \( M \).

Math­em­aticians in gen­er­al value proofs of new facts much more highly than el­eg­ant new proofs of old res­ults. It is worth com­ment­ing why [9] is an ex­cep­tion to this rule. The earli­er proofs of the Gen­er­al­ized Gauss–Bon­net The­or­em were vir­tu­ally a dead end while, as we shall see be­low, Chern’s in­trins­ic proof was a key that opened the door to the secrets of char­ac­ter­ist­ic classes.

Characteristic classes

The coframe bundle, \( F (M ) \), that keeps re­appear­ing in our story, is an im­port­ant ex­ample of a math­em­at­ic­al struc­ture known as a prin­cip­al \( G \)-bundle. These were first defined and their study be­gun only in the late 1930s, but their im­port­ance was quickly re­cog­nized by to­po­lo­gists and geo­met­ers, and the the­ory un­der­went in­tens­ive de­vel­op­ment dur­ing the 1940s. By the end of that dec­ade the beau­ti­ful clas­si­fic­a­tion the­ory had been worked out, and with it the re­lated the­ory of “char­ac­ter­ist­ic classes”, a concept whose im­port­ance for the math­em­at­ics of the lat­ter half of the twen­ti­eth cen­tury it would be dif­fi­cult to ex­ag­ger­ate. (As we will see be­low, in the lan­guage we have been us­ing, the clas­si­fic­a­tion prob­lem is the equi­val­ence prob­lem for prin­cip­al bundles, and char­ac­ter­ist­ic classes are in­vari­ants for this equi­val­ence prob­lem).

In or­der to ex­plain Chern’s role in these im­port­ant de­vel­op­ments we will first re­view some of the ba­sic math­em­at­ic­al back­ground of the the­ory.

We will con­sider only the case of a Lie group \( G \). Since the the­ory is es­sen­tially the same for a Lie group and one of its max­im­al com­pact sub­groups, we will also as­sume that \( G \) is com­pact. A “space” will mean a para­com­pact to­po­lo­gic­al space, and a \( G \)-space will mean a space, \( P \), to­geth­er with a con­tinu­ous right ac­tion of \( G \) on \( P \). We will write \( R_g \) for the homeo­morph­ism \( p \mapsto pg \). The \( G \)-space \( P \) is called a prin­cip­al \( G \)-bundle if the ac­tion is free, i.e., if for all \( p \) in \( P \), \( R_g (p) \neq p \) un­less \( g \) is the iden­tity ele­ment \( e \) of \( G \). More spe­cific­ally, \( P \) is called a prin­cip­al \( G \)-bundle over a space \( X \) if we are giv­en some fixed homeo­morph­ism of \( X \) with the or­bit space \( P/G \), or equi­val­ently if there is giv­en a “pro­jec­tion map” \[ \Pi : P \rightarrow X \] such that the \( G \) or­bits of \( P \) are ex­actly the “fibers” \( \Pi^{-1} (x) \) of the map \( \Pi \). \( P \) is called the total space of the bundle, and we of­ten de­note the bundle by the same sym­bol as the total space. A map \( \sigma : X \rightarrow P \) that is a left in­verse to \( \Pi \) is called a sec­tion. Two \( G \)-bundles over \( X \), \( \Pi_i : P_i \rightarrow X \), \( i = 1, 2 \) are con­sidered “equi­val­ent” if there is a \( G \)-equivari­ant homeo­morph­ism \[ \varphi : P_1 \rightarrow P_2 \quad\text{such that}\quad \Pi_1 = \Pi_2 \circ \varphi .\] The prin­cip­al \( G \)-bundle over \( X \) defined by \( P = X \times G \) with \( R_g (x, \gamma) = (x, \gamma g) \) and \( \Pi(x, \gamma) = x \) is called the product bundle, and any bundle equi­val­ent to the product bundle is called a trivi­al bundle. Clearly \( x \mapsto (x, e) \) is a sec­tion of the product bundle, so any trivi­al bundle has a sec­tion. Con­versely, if \( \Pi : P \rightarrow X \) has a sec­tion \( \sigma \), then \( \varphi(x, g) = R_g (\sigma(x)) \) is an equi­val­ence of the product bundle with \( P \), i.e., a prin­cip­al \( G \)-bundle is trivi­al if and only if it ad­mits a sec­tion. We will de­note the set of equi­val­ence classes [P ] of prin­cip­al \( G \)-bundles \( P \) over \( X \) by \( \operatorname{Bndl}_G (X) \).

Giv­en a prin­cip­al \( G \)-bundle \( \Pi : P \rightarrow X \) and a con­tinu­ous map \( f : Y \rightarrow X \), we can define a bundle \( f^\ast (P ) \) over \( Y \), called the bundle in­duced from \( P \) by the map \( f \). Its total space is \[ \bigl\{(p, y) \in P \times Y \bigm| \Pi(p) = f (y)\bigr\} ,\] with the pro­jec­tion \( (p, y) \mapsto y \) and the \( G \)-ac­tion \[ R_g (p, y) = \bigl(R_g (p), y\bigr) .\] It is easy to see that \( f^\ast \) maps equi­val­ent bundles to equi­val­ent bundles, so it in­duces a map (also de­noted by \( f^\ast \)) from \( \operatorname{Bndl}_G (X) \) to \( \operatorname{Bndl}_G (Y ) \). If \( \Pi : P \rightarrow X \) is a prin­cip­al \( G \)-bundle then \( \Pi^\ast (P ) \) is a prin­cip­al \( G \)-bundle over the total space \( P \), called the “square” of the ori­gin­al bundle. In fact this bundle is al­ways trivi­al, since it ad­mits the “di­ag­on­al” sec­tion \( p \mapsto (p, p) \). As we will see be­low, this simple ob­ser­va­tion is the secret be­hind trans­gres­sion!

The first non­trivi­al fact in the the­ory is the so-called “cov­er­ing ho­mo­topy the­or­em”; it says that the in­duced map \[ f^\ast : \operatorname{Bndl}_G (X) \rightarrow \operatorname{Bndl}_G (Y ) \] de­pends only on the ho­mo­topy class \( [f ] \) of \( f \). We can para­phrase this by say­ing that \( \operatorname{Bndl}_G (\,\cdot\,) \) is a con­trav­ari­ant func­tor from the cat­egory of spaces and ho­mo­topy classes of maps to the cat­egory of sets. Now a co­homo­logy the­ory is also such a func­tor, and a char­ac­ter­ist­ic class for \( G \)-bundles can be defined as simply a nat­ur­al trans­form­a­tion from \( \operatorname{Bndl}_G (\,\cdot\,) \) to some co­homo­logy the­ory \( H^\ast (\,\cdot\,) \). Of course this fancy lan­guage isn’t es­sen­tial and was only in­ven­ted about the same time as bundle the­ory. It just says that a char­ac­ter­ist­ic class \( c \) is a func­tion that as­signs to each prin­cip­al \( G \)-bundle \( P \) over any space \( X \) an ele­ment \( c(P ) \) in \( H^\ast (X) \), with the “nat­ur­al­ity” prop­erty that \[ c\bigl(f^\ast (P )\bigr) = f^\ast \bigl(c(P )\bigr) ,\] for any con­tinu­ous \( f : Y \rightarrow X \). We fix some co­homo­logy the­ory \( H^\ast (\,\cdot\,) \) and de­note by \( \operatorname{Char}(G) \) the set of all char­ac­ter­ist­ic classes for \( G \)-bundles. Since \( H^\ast (X) \) has the struc­ture of a ring with unit, so does \( \operatorname{Char}(G) \), and the char­ac­ter­ist­ic class prob­lem for \( G \) is the prob­lem of ex­pli­citly identi­fy­ing this ring. Note that a trivi­al bundle is in­duced from a map to a space with one point, so all its char­ac­ter­ist­ic classes (ex­cept the unit class) must be zero. More gen­er­ally, equal­ity of all char­ac­ter­ist­ic classes of a bundle is a ne­ces­sary (and in some cir­cum­stances suf­fi­cient) test for their equi­val­ence, and this is one of the im­port­ant uses of char­ac­ter­ist­ic classes.

The re­mark­able and beau­ti­ful clas­si­fic­a­tion the­or­em for prin­cip­al \( G \)-bundles “solves” the clas­si­fic­a­tion prob­lem at least in the sense of re­du­cing it to a stand­ard prob­lem of ho­mo­topy the­ory. Giv­en spaces \( X \) and \( Z \) let \( [X, Z] \) de­note the set of ho­mo­topy classes of maps of \( X \) in­to \( Z \). Note that \( [\,\cdot\, , Z] \) is a con­trav­ari­ant func­tor, much like \( \operatorname{Bndl}_G \) — any map \( f : Y \rightarrow X \) in­duces a pull-back map \[ f^\ast : [h] \mapsto [h \circ f ] \] of \( [X, Z] \) to \( [Y, Z] \). Moreover if \( \Pi : P \rightarrow Z \) is any prin­ciple \( G \)-bundle then we have a map \[ [h] \mapsto [h^\ast (P)] \] of \( [X, Z] \) to \( \operatorname{Bndl}_G (X) \) that is “nat­ur­al” (i.e., it com­mutes with all “pull-back” maps \( f^\ast \)). We call \( P \) a uni­ver­sal prin­cip­al \( G \)-bundle if the lat­ter map is biject­ive. The heart of the clas­si­fic­a­tion the­or­em is the fact that uni­ver­sal \( G \)-bundles do ex­ist. In fact it can be shown that a prin­cip­al \( G \)-bundle is uni­ver­sal provided its total space is con­tract­ible, and there are even a num­ber of meth­ods for ex­pli­citly con­struct­ing such bundles.

We will de­note by \( \mathcal{U}_G \) some choice of uni­ver­sal prin­cip­al \( G \)-bundle. Its base space will be de­noted by \( \mathcal{B}_G \) and is called the clas­si­fy­ing space for \( G \). (Al­though \( \mathcal{B}_G \) is not unique, its ho­mo­topy type is). If \( \Pi : P \rightarrow X \) is any prin­cip­al \( G \)-bundle then, by defin­i­tion of uni­ver­sal, there is a unique ho­mo­topy class \( [h] \) of maps of \( X \) to \( \mathcal{B}_G \) such that \( P \) is equi­val­ent to \( h^\ast (\mathcal{U}_G ) \). Any rep­res­ent­at­ive \( h \) is called a clas­si­fy­ing map for \( P \). Clearly if \( f : Y \rightarrow X \) then \( h \circ f \) is a clas­si­fy­ing map for \( f^\ast (P ) \). Also, the clas­si­fy­ing map for \( \mathcal{U}_G \) is just the iden­tity map of \( \mathcal{B}_G \).

It is now easy to give a solu­tion of sorts to the char­ac­ter­ist­ic class prob­lem for \( G \); namely \( \operatorname{Char}(G) \) is ca­non­ic­ally iso­morph­ic to \( H^\ast (\mathcal{B}_G ) \). In fact each \( c \in H^\ast (\mathcal{B}_G ) \) defines a char­ac­ter­ist­ic class (also de­noted by \( c \)) by the for­mula \[ c(P ) = f^\ast (c) ,\] where \( f \) is a clas­si­fy­ing map for \( P \), and the in­verse map is just \( c \mapsto c(\mathcal{U}_G ) \).

This is a dis­til­la­tion of ideas de­veloped between 1935 and 1950 by Chern, Ehresmann, Hopf, Feld­bau, Pontry­agin, Steen­rod, Stiefel, and Whit­ney. While el­eg­ant in its sim­pli­city, the above ver­sion is still too ab­stract and gen­er­al to be of use in find­ing \( \operatorname{Char}(G) \) for a spe­cif­ic group \( G \). It is also of little use in cal­cu­lat­ing the char­ac­ter­ist­ic classes of bundles that come up in geo­met­ric prob­lems, for it is not of­ten an easy mat­ter to find a clas­si­fy­ing map from geo­met­ric data. We shall dis­cuss how Chern put flesh on these bones by find­ing con­crete mod­els for clas­si­fy­ing spaces and, more im­port­antly, by show­ing how to cal­cu­late ex­pli­citly de Rham the­ory rep­res­ent­at­ives of many char­ac­ter­ist­ic classes from the curvature forms of con­nec­tions.

Let \( \mathbf{V}(n, N + n) \) de­note the Stiefel man­i­fold of \( n \)-frames in \( \mathbf{R}^{N +n} \), con­sist­ing of all or­thonor­mal se­quences \[ e = (e_1 , \dots, e_n) \] of vec­tors in \( \mathbf{R}^{N +n} \). There is an ob­vi­ous free ac­tion of \( \mathbf{O}(n) \) on \( \mathbf{V}(n, N + n) \), and the or­bit of \( e \) con­sists of all \( n \)-frames span­ning the same \( n \)-di­men­sion­al lin­ear sub­space that \( e \) does. Thus we have an \( \mathbf{O}(n) \) prin­cip­al bundle \[ \Pi : \mathbf{V}(n, N +n) \rightarrow \mathbf{Gr}(n, N +n) ,\] where \( \mathbf{Gr}(n, N +n) \) is the Grass­man­ni­an of all \( n \)-di­men­sion­al lin­ear sub­spaces of \( \mathbf{R}^{N +n} \). In the early 1940s it was known from res­ults of Steen­rod and Whit­ney that this bundle is “uni­ver­sal for com­pact k-di­men­sion­al poly­hedra”, provided \( N \geq k + 1 \). This means that for any com­pact poly­hed­ral space \( X \), with \( \dim(X) \leq k \), every prin­cip­al \( \mathbf{O}(n) \) bundle over \( X \) is of the form \[ h^\ast \bigl(\mathbf{V}(n, N +n)\bigr) \] for a unique \( [h] \) in \( [X, \mathbf{Gr}(n, N + n)] \). In [12] Chern and Y. F. Sun gen­er­al­ized these res­ults to show that this bundle is also uni­ver­sal for com­pact \( k \)-di­men­sion­al ANRs. (If one wants uni­ver­sal bundles in the strict sense de­scribed above, one need only form the ob­vi­ous in­duct­ive lim­it, \[ \Pi : \mathbf{V}(n, \infty) \rightarrow \mathbf{Gr}(n, \infty) ,\] by let­ting \( N \) tend to in­fin­ity. But for the fi­nite di­men­sion­al prob­lems of geo­metry it is prefer­able to stick with these fi­nite di­men­sion­al mod­els). By re­pla­cing the real num­bers re­spect­ively by the com­plex num­bers and the qua­ternions, Chern and Sun proved ana­log­ous res­ults for the oth­er clas­sic­al groups \( \mathbf{U}(n) \) and \( \mathbf{Sp}(n) \). They went on to note that if \( G \) is any com­pact Lie group, then by tak­ing a faith­ful rep­res­ent­a­tion of \( G \) in some \( \mathbf{O}(n) \), \( \mathbf{V}(n, N +n) \) be­comes a prin­cip­al \( G \) bundle by re­stric­tion, and the cor­res­pond­ing or­bit space \[ \mathbf{V}(n, N + n)/G \] be­comes a clas­si­fy­ing space \( \mathcal{B}_G \) for com­pact ANRs of di­men­sion \( \leq k \).

The Grass­man­ni­ans make good mod­els for clas­si­fy­ing spaces, for they are well-stud­ied ex­pli­cit ob­jects whose co­homo­logy can be in­vest­ig­ated us­ing both al­geb­ra­ic and geo­met­ric tech­niques. From such com­pu­ta­tions Chern knew that there was an n-di­men­sion­al “Euler class” \( e \) in \( \operatorname{Char}(\mathbf{SO}(n)) \). If \( M \) is a smooth, com­pact, ori­ented \( n \)-di­men­sion­al man­i­fold then \( e(F (M )) \in H^n (M ) \) when eval­u­ated on the fun­da­ment­al class of \( M \) is just \( \chi(M ) \). One can thus in­ter­pret the Gen­er­al­ized Gauss–Bon­net The­or­em as say­ing that \( \lambda = \operatorname{Pf}(\Psi) \) rep­res­ents \( e(F (M )) \) in de Rham co­homo­logy. This in­spired Chern to look for a gen­er­al tech­nique for rep­res­ent­ing char­ac­ter­ist­ic classes by de Rham classes. This was in 1944–1945, while Chern was in Prin­ceton, and he dis­cussed this prob­lem fre­quently with his friend An­dré Weil who en­cour­aged him in this search.

It might seem nat­ur­al to start by try­ing to rep­res­ent \( \mathbf{SO}(n) \) char­ac­ter­ist­ic classes by closed dif­fer­en­tial forms, but Chern made what was to be a cru­cial ob­ser­va­tion: the co­homo­logy of the real Grass­man­ni­ans is com­plic­ated. In par­tic­u­lar it con­tains a lot of \( \mathbf{Z}_2 \) tor­sion, and this part of the co­homo­logy is in­vis­ible to de Rham the­ory. On the oth­er hand Chern knew that Ehresmann, in his thes­is, had cal­cu­lated the ho­mo­logy of com­plex Grass­man­ni­ans and showed there was no tor­sion. In fact Ehresmann showed that cer­tain ex­pli­cit al­geb­ra­ic cycles (the “Schubert cells”) form a free basis for the ho­mo­logy over \( \mathbf{Z} \). It fol­lows from de Rham’s The­or­em that all the co­homo­logy classes for \( \mathcal{B}_{\mathbf{U}(n)} \) can be rep­res­en­ted by closed dif­fer­en­tial forms. These forms, when pulled back by the clas­si­fy­ing map of a prin­cip­al \( \mathbf{U}(n) \)-bundle, will then rep­res­ent the char­ac­ter­ist­ic classes of the bundle in de Rham co­homo­logy. While this is fine in the­ory, it still de­pends on know­ing a clas­si­fy­ing map, while what is needed in prac­tice is a meth­od to cal­cu­late these char­ac­ter­ist­ic forms from geo­met­ric data. We now ex­plain Chern’s beau­ti­ful al­gorithm for do­ing this.

Let \( \Pi : P \rightarrow M \) be a smooth prin­ciple \( \mathbf{U}(n) \)-bundle over a smooth man­i­fold \( M \). Re­call that a con­nec­tion for \( P \) can be re­garded as a 1-form \( \omega \) on \( P \) with val­ues in the Lie al­gebra of \( \mathbf{U}(n) \), \( L(\mathbf{U}(n)) \), which con­sists of all \( n {\times} n \) skew-Her­mitian com­plex matrices. Equi­val­ently we can re­gard \( \omega \) as an \( n {\times} n \) mat­rix of com­plex-val­ued 1-forms \( \omega_{ij} \) on \( P \) sat­is­fy­ing \( \omega_{ji} = -\bar{\omega}_{ij} \), and sim­il­arly for the as­so­ci­ated curvature 2-forms \( \Omega_{ij} \).

We will de­note by \( \mathcal{R} \) the ring of com­plex-val­ued poly­no­mi­al func­tions on the vec­tor space \( L(\mathbf{U}(n)) \). Us­ing the usu­al basis for the \( L(\mathbf{U}(n)) \), we can identi­fy \( \mathcal{R} \) with com­plex poly­no­mi­als in the \( 2n(n - 1) \) vari­ables \( X_{ij} \), \( Y_{ij} \), \( 1 \leq i < j \leq n \) and the \( n \) vari­ables \( Y_{ii} \), \( 1 \leq i \leq n \). \( Z \) will de­note the \( n {\times} n \) mat­rix of ele­ments in \( \mathcal{R} \) defined by \begin{align*} & Z_{ij} = X_{ij} + \sqrt{-1}\,Y_{ij},\\ & Z_{ji} = -X_{ij} + \sqrt{-1}\,Y_{ij}\quad\text{and}\\ & Z_{ii} = \sqrt{-1}\,Y_{ii} \end{align*} for \( 1 \leq i < j \leq n \). We can also re­gard \( Z \) as rep­res­ent­ing the gen­er­al ele­ment of \( L(\mathbf{U}(n)) \), and we will write \( Q(Z) \) rather than \( Q(X_{ij} , Y_{ij}) \) to de­note ele­ments of \( \mathcal{R} \). The ad­joint ac­tion of the group \( \mathbf{U}(n) \) on its Lie al­gebra \( L(\mathbf{U}(n)) \) is now giv­en by \[ \operatorname{ad}(g)(Z) = gZg^{-1} ,\] just as in the \( \mathbf{SO}(n) \) case above, and as in that case we define the ad­joint ac­tion of \( \mathbf{U}(n) \) on \( \mathcal{R} \) by \[ \bigl(\operatorname{ad}(g)Q\bigr)(Z) = Q\bigl(\operatorname{ad}(g)Z\bigr) .\] As be­fore we de­note by \( \mathcal{R}^{\operatorname{ad}} \) the sub­ring of \( \mathcal{R} \) con­sist­ing of ad in­vari­ant poly­no­mi­als. Once again we can sub­sti­tute the curvature forms \( \Omega_{ij} \) for the \( Z_{ij} \) in an ele­ment \( Q(Z) \) in \( \mathcal{R} \), and ob­tain a dif­fer­en­tial form \( Q(\Omega) \) on \( P \); if \( Q \) is ho­mo­gen­eous of de­gree \( d \) in its vari­ables then \( Q(\Omega) \) is a \( 2d \)-form. The same ar­gu­ment as in the \( \mathbf{SO}(n) \) case shows that if \( Q \in \mathcal{R}^{\operatorname{ad}} \) then \( Q(\Omega) \) is the pull-back of a uniquely de­term­ined form \( Q(\Psi) \) on \( M \). Us­ing the Bi­an­chi iden­tity, Chern showed that \[ \mathit{dQ}(\Psi) = 0 \] (cf. [e2], p. 297), so \( Q(\Psi) \) rep­res­ents an ele­ment \( [Q(\Psi)] \) in \( H^\ast (M ) \), the com­plex de Rham co­homo­logy ring of \( M \). If we use a dif­fer­ent con­nec­tion \( \omega \) on \( P \) with curvature mat­rix \( \Omega \) then we get a dif­fer­ent closed form \( Q(\Psi ) \) on \( M \) with \[ \Pi^\ast \bigl(Q(\Psi^{\prime})\bigr) = Q(\Omega^{\prime}) .\] What is the re­la­tion between \( Q(\Psi^{\prime} ) \) and \( Q(\Psi) \)? Weil provided Chern with the ne­ces­sary lemma: they dif­fer by an ex­act form, so that \( [Q(\Psi)] \) is a well-defined ele­ment of \( H^\ast (M ) \), in­de­pend­ent of the con­nec­tion. We will de­note it by \( \hat{Q}(P ) \). (Weil’s lemma can be de­rived as a co­rol­lary of the fact that \( Q(\Psi) \) is closed. For the easy but clev­er proof see [e2], p. 297).

If \( h : M \rightarrow M \) is a smooth map, then a con­nec­tion on \( P \) “pulls-back” nat­ur­ally to one on the \( \mathbf{U}(n) \)-bundle \( h^\ast (P ) \) over \( M \). The curvature forms like­wise are pull-backs, from which it is im­me­di­ate that \[ Q\bigl(h^\ast (P )\bigr) = h^\ast \bigl(Q(P )\bigr) .\] In oth­er words, \( Q \mapsto \hat{Q} \) is a map from \( \mathcal{R}^{\operatorname{ad}} \) in­to \( \operatorname{Char}(\mathbf{U}(n)) \). It is clearly a ring ho­mo­morph­ism, and in re­cog­ni­tion of Weil’s lemma Chern called it the Weil ho­mo­morph­ism, but it is more com­monly re­ferred to as the Chern–Weil ho­mo­morph­ism.

For \( \mathbf{U}(n) \) the ring \( \mathcal{R}^{\operatorname{ad}} \) of ad-in­vari­ant poly­no­mi­als on its Lie al­gebra has an el­eg­ant and ex­pli­cit de­scrip­tion that fol­lows eas­ily from the di­ag­on­al­iz­ab­il­ity of skew-Her­mitian op­er­at­ors and the clas­sic clas­si­fic­a­tion of sym­met­ric poly­no­mi­als. Ex­tend the ad­joint ac­tion of \( \mathbf{U}(n) \) to the poly­no­mi­al ring \( \mathcal{R}[t] \) by let­ting it act trivi­ally on the new in­de­term­in­ate \( t \). The char­ac­ter­ist­ic poly­no­mi­al \[ \operatorname{det}(Z + tI) = \sum_{k=0}^n \sigma_k (Z)\,t^{n-k} \] is clearly ad-in­vari­ant, and hence its coef­fi­cients \( \sigma_k (Z) \) be­long to \( \mathcal{R}^{\operatorname{Ad}} \). Sub­sti­tut­ing a par­tic­u­lar mat­rix for \( Z \) in \( \sigma_k (Z) \) gives the \( k \)-th ele­ment­ary sym­met­ric func­tion of its ei­gen­val­ues; in par­tic­u­lar \( \sigma_1 (Z) = \operatorname{trace}(Z) \) and \( \sigma_n (Z) = \det(Z) \). Now if \[ P (t_1, \dots, t_n ) \in \mathbf{C}[t_1 , \dots, t_n ] \] then of course \[ P \bigr(\sigma_1 [Z],\dots, \sigma_n [Z]\bigl) \] is also in \( \mathcal{R}^{\operatorname{ad}} \). In fact, \[ \mathcal{R}^{\operatorname{ad}} = C[\sigma_1, \dots,\sigma_n ] ,\] i.e., \[ P (t_1 , \dots, t_n ) \mapsto P \bigl(\sigma_1 [Z], \dots, \sigma_n [Z]\bigr) \] is a ring iso­morph­ism. From this fact, to­geth­er with Ehresmann’s ex­pli­cit de­scrip­tion of the ho­mo­logy of com­plex Grass­man­ni­ans, Chern was eas­ily able to veri­fy that the Chern–Weil ho­mo­morph­ism is in fact an iso­morph­ism of \( \mathcal{R}^{\operatorname{ad}} \) with \( \operatorname{Char}(\mathbf{U}(n)) \). For tech­nic­al reas­ons it is con­veni­ent to renor­mal­ize the poly­no­mi­als \( \sigma_k (Z) \), de­fin­ing \[ \gamma_k (Z) = \sigma_k \Bigl( \frac{1}{2\pi i} Z\Bigr) .\] Then we get a \( \mathbf{U}(n) \)-char­ac­ter­ist­ic class \( c_k = \hat{\gamma}_k \) of di­men­sion \( 2k \), called the \( k \)-th Chern class, and these \( n \) classes \( c_1 , \dots \), \( c_n \) are poly­no­mi­al gen­er­at­ors for the char­ac­ter­ist­ic ring \( \operatorname{Char}(\mathbf{U}(n)) \); that is each \( \mathbf{U}(n) \)-char­ac­ter­ist­ic class \( c \) can be writ­ten uniquely as a poly­no­mi­al in the Chern classes.

If \( F (Z) \) is a form­al power series, \( F = \sum^\infty_0 F_r \), where \( F_r \) is a ho­mo­gen­eous poly­no­mi­al of de­gree \( r \), then for fi­nite di­men­sion­al spaces, \( \hat{F}_r \) will van­ish for large \( r \) so \[ \hat{F} = \sum^\infty_0\hat{F}_r \] will be a well-defined char­ac­ter­ist­ic class. Many im­port­ant classes were defined in this way by Hirzebruch, and Chern used the power series \[ E(Z) = \operatorname{trace}\Bigl(\exp\Bigl( \frac{1}{2\pi i} Z\Bigr)\Bigr) \] to define the Chern char­ac­ter, \[ \mathbf{ch} = \hat{E} .\] It plays a vi­tal role in the Atiyah–Sing­er In­dex The­or­em.

Chern also de­veloped a gen­er­al­iz­a­tion of the Chern–Weil ho­mo­morph­ism for an ar­bit­rary com­pact Lie group \( G \). The ad­joint ac­tion of \( G \) on its Lie al­gebra \( L(G) \) in­duces one on the ring \( \mathcal{R} \) of com­plex-val­ued poly­no­mi­al func­tions on \( L(G) \), so we have a sub­ring \( \mathcal{R}^{\operatorname{ad}} \) of ad­joint in­vari­ant poly­no­mi­als. Sub­sti­tut­ing curvature forms of \( G \)-con­nec­tions on \( G \)-prin­cip­al bundles in­to such in­vari­ant poly­no­mi­als \( Q \), we get as above a Chern–Weil ho­mo­morph­ism \( Q \mapsto Q \) of \( \mathcal{R}^{\operatorname{ad}} \) to the char­ac­ter­ist­ic ring \( \operatorname{Char}(G) \) (defined with re­spect to com­plex de Rham co­homo­logy) and this is again an iso­morph­ism. Of course, for gen­er­al \( G \) the ho­mo­logy of the clas­si­fy­ing space \( \mathcal{B}_G \) will have tor­sion, so there will be oth­er char­ac­ter­ist­ic classes bey­ond those picked up by de Rham the­ory. Moreover the ex­pli­cit de­scrip­tion of the ring of ad­joint in­vari­ant poly­no­mi­als is in gen­er­al fairly com­plic­ated.

Chern left the sub­ject of char­ac­ter­ist­ic classes for nearly twenty years, but then re­turned to it in 1974 in a now fam­ous joint pa­per with J. Si­mons [21]. This pa­per is a de­tailed and el­eg­ant study of the phe­nomen­on of trans­gres­sion in prin­cip­al bundles. Let \( M \) be an \( n \)-di­men­sion­al smooth man­i­fold, \( \Pi : P \rightarrow M \) a smooth prin­cip­al \( G \)-bundle over \( M \), \( \omega \) a \( G \)-con­nec­tion in \( P \), and \( \Omega \) the mat­rix of curvature 2-forms. Giv­en an ad­joint in­vari­ant poly­no­mi­al \( Q \) on \( L(G) \), ho­mo­gen­eous of de­gree \( \ell \), we have a glob­ally defined closed \( 2\ell \)-form \( Q(\Psi) \) on \( M \) that rep­res­ents the char­ac­ter­ist­ic class \( \hat{Q}(P ) \in H^{2\ell} (M ) \), and that is char­ac­ter­ized by \[ \Pi^{\ast} \bigl(Q(\Psi)\bigr) = Q(\Omega) .\] Chern and Si­mons first point out the simple reas­on why \( Q(\Omega) \) must be an ex­act form on \( P \). In­deed, by the nat­ur­al­ity of char­ac­ter­ist­ic classes un­der pull-back, \( Q(\Omega) \) rep­res­ents \( \hat{Q}(\Pi^{\ast} (P )) \). But as we saw earli­er, \( \Pi^{\ast} (P ) \), the “square” of the bundle \( P \), is a prin­cip­al \( G \)-bundle over \( P \) with a glob­al cross-sec­tion, hence it is trivi­al and all of its char­ac­ter­ist­ic classes must van­ish. In par­tic­u­lar \[ \hat{Q}\bigl(\Pi^{\ast} (P )\bigr) = 0 ,\] i.e., \( Q(\Omega) \) is ex­act.

They next write down an ex­pli­cit for­mula in terms of \( Q \), \( \omega \), and \( \Omega \) for a \( (2 \ell- 1) \)-form \( \mathit{TQ}(\omega) \) on \( P \), and show that \[ d\mathit{TQ}(\omega) = Q(\Omega) .\] \( \mathit{TQ}(\omega) \) is nat­ur­al un­der pull-back of a bundle and its con­nec­tion. Now sup­pose \( 2 > n \). Then \( Q(\Psi) = 0 \), so of course \( Q(\Omega) = 0 \), i.e., in this case \( \mathit{TQ}(\omega) \) is closed, and so defines an ele­ment \( [\mathit{TQ}(\omega)] \) of \( H^{2\ell -1} (P ) \). If \( 2\ell > n + 1 \) Chern and Si­mons show this co­homo­logy class is in­de­pend­ent of the choice of con­nec­tion \( \omega \), and so defines a “sec­ond­ary char­ac­ter­ist­ic class”. However if \( 2 = n + 1 \) then they show that \( [\mathit{TQ}(\omega)] \) does de­pend on the choice of con­nec­tion \( \omega \).

They now con­sider the case \( G =\mathbf{GL}(n, \mathbf{R}) \) and con­sider the ad­joint in­vari­ant \( n \) poly­no­mi­als \( Q_k \) defined by \[ \det(X + tI) =\sum^n_{i=0} Q_i (X)\,t^{n-i} .\] Tak­ing \( Q = Q_{2k-1} \) they again show \( Q(\Omega) = 0 \) provided \( \omega \) re­stricts to an \( \mathbf{O}(n) \) con­nec­tion on an \( \mathbf{O}(n) \)-sub­bundle of \( P \), so of course in this case too we have a co­homo­logy class \( [\mathit{TQ}(\omega)] \). They spe­cial­ize to the case that \( P \) is the bundle of bases for the tan­gent bundle of \( M \) and \( \omega \) is the Levi-Civ­ita con­nec­tion of a Rieman­ni­an struc­ture. Then \( [\mathit{TQ}(\omega)] \) is defined, but de­pends in gen­er­al on the choice of Rieman­ni­an met­ric. Now they prove a re­mark­able and beau­ti­ful fact — \( [\mathit{TQ}(\omega)] \) is in­vari­ant un­der con­form­al changes of the Rieman­ni­an met­ric! Such con­form­al in­vari­ants have re­cently been ad­op­ted by phys­i­cists in for­mu­lat­ing so-called con­form­al quantum field the­or­ies.

Chern also re­turned to the con­sid­er­a­tion of char­ac­ter­ist­ic classes and trans­gres­sion in an­oth­er joint pa­per, this one with R. Bott [19]. Here they con­sider holo­morph­ic bundles over com­plex ana­lyt­ic man­i­folds, where there is a re­fined ex­ter­i­or cal­cu­lus, us­ing the \( \partial \) and \( \bar{\partial} \) op­er­at­ors, and they prove a trans­gres­sion for­mula for the top Chern form of a Her­mitian struc­ture with re­spect to the op­er­at­or \( i\partial \bar{\partial} \). This work has ap­plic­a­tions both to com­plex geo­metry (es­pe­cially the study of the zer­os of holo­morph­ic sec­tions), and to al­geb­ra­ic num­ber the­ory. In re­cent years it has played an im­port­ant role in pa­pers by J. M. Bis­mut, H. Gil­let, and C. Soulé.

“Retirement”

For most math­em­aticians, re­tire­ment is a one-time event fol­lowed by a peri­od of de­clin­ing math­em­at­ic­al activ­ity. But as with so much else, Chern’s at­ti­tude to­wards re­tire­ment is highly non­stand­ard. Both au­thors re­mem­ber well at­tend­ing a series of en­joy­able so-called re­tire­ment parties for Chern, as he re­tired first from UC Berke­ley, then sev­er­al years later as Dir­ect­or of MSRI, etc. But in each case, in­stead of re­tir­ing, Chern merely re­placed one de­mand­ing job with an­oth­er.

Fi­nally, in 1992, Dr. Hu Guo-Ding took over as dir­ect­or of the Nankai In­sti­tute of Math­em­at­ics and Chern de­clared him­self truly re­tired. In fact though, he travels back to Nankai one or more times each year and con­tin­ues to play an act­ive role in the life of the In­sti­tute. The In­sti­tute now has an ex­cel­lent lib­rary, has be­come in­creas­ingly act­ive in in­ter­na­tion­al ex­changes, and has many well-trained young­er mem­bers. In 1995, the oc­ca­sion of the tenth an­niversary of the Nankai In­sti­tute was cel­eb­rated with a highly suc­cess­ful in­ter­na­tion­al con­fer­ence, at­ten­ded by many well-known phys­i­cists and math­em­aticians.

Chern also con­tin­ues to be very act­ive in math­em­at­ic­al re­search, and when asked why he doesn’t slow down and take it a little easi­er, his stock “ex­cuse” is that he does not know how to do something else. He says he tries to work in areas that he feels have a fu­ture, avoid­ing the cur­rent fash­ions. His re­cent in­terests have been Lie sphere geo­metry, sev­er­al com­plex vari­ables, and par­tic­u­larly Finsler geo­metry. Chern’s in­terest in the lat­ter sub­ject has a long his­tory. Already in 1948 he solved the equi­val­ence prob­lem for the sub­ject in “Loc­al Equi­val­ence and Eu­c­lidean con­nec­tions in Finsler spaces” (re­prin­ted in [28]). Chern feels that the time is now ripe to re­cast all the beau­ti­ful glob­al res­ults of Rieman­ni­an geo­metry of the past sev­er­al dec­ades in the Finsler con­text, and he points out that think­ing of Rieman­ni­an geo­metry as a spe­cial case of Finsler geo­metry was already ad­voc­ated by Dav­id Hil­bert in his twenty-third prob­lem at the turn of the last cen­tury. Chern him­self has re­cently taken some steps in that dir­ec­tion, in “On Finsler geo­metry” (C. R. Acad. Sci. Par­is, t. 314, Série I, p. 757–761, 1992), and with Dav­id Bao, “On a not­able con­nec­tion in Finsler geo­metry” (Hou­s­ton Journ­al of Math., v. 19, no. 1, 1993). He has also re­cently spelled out the gen­er­al pro­gram in a pa­per that is as yet un­pub­lished, “Rieman­ni­an geo­metry as a spe­cial case of Finsler geo­metry”.

Works

[1] article S.-S. Chern: “Sur la géométrie d’une équa­tion différen­ti­elle du troisième or­dre” [On the geo­metry of a dif­fer­en­tial equa­tion of third or­der], C. R. Acad. Sci., Par­is 204 (1937), pp. 1227–​1229. JFM 63.​0419.​01 Zbl 0016.​16401

[2] article S.-S. Chern: “Sur la géométrie d’un sys­tème d’équa­tions différen­ti­elles du second or­dre” [On the geo­met­ery of a sys­tem of second or­der dif­fer­en­tial equa­tions], Bull. Sci. Math., II. Ser. 63 (1939), pp. 206–​212. MR 0000889 Zbl 0023.​07701

[3] article S.-S. Chern: “The geo­metry of high­er path-spaces,” J. Chin. Math. Soc. 2 (1940), pp. 247–​276. MR 0004531 Zbl 0063.​00828

[4] article S.-S. Chern: “The geo­metry of the dif­fer­en­tial equa­tion \( y^{\prime\prime\prime}=F(x,y,y^{\prime},y^{\prime\prime}) \),” Sci. Rep. Nat. Tsing Hua Univ. (A) 4 (1940), pp. 97–​111. MR 0004538 Zbl 0024.​19801

[5] article S.-S. Chern: “On in­teg­ral geo­metry in Klein spaces,” Ann. Math. (2) 43 : 1 (January 1942), pp. 178–​189. MR 0006075 Zbl 0147.​22303

[6] article S.-S. Chern: “The geo­metry of iso­trop­ic sur­faces,” Ann. Math. (2) 43 : 3 (July 1942), pp. 545–​559. MR 0006477 Zbl 0060.​39207

[7] article S.-S. Chern: “On a Weyl geo­metry defined from an \( (n-1) \)-para­met­er fam­ily of hy­per­sur­faces in a space of \( n \) di­men­sions,” Sci. Rec. 1 (1942), pp. 7–​10. MR 0007637 Zbl 0060.​39208

[8] article S.-S. Chern: “A gen­er­al­iz­a­tion of the pro­ject­ive geo­metry of lin­ear spaces,” Proc. Natl. Acad. Sci. U. S. A. 29 : 1 (January 1943), pp. 38–​43. MR 0008192 Zbl 0060.​39209

[9] article S.-S. Chern: “A simple in­trins­ic proof of the Gauss–Bon­net for­mula for closed Rieman­ni­an man­i­folds,” Ann. Math. (2) 45 : 4 (October 1944), pp. 747–​752. MR 0011027 Zbl 0060.​38103

[10] article S.-S. Chern: “Char­ac­ter­ist­ic classes of Her­mitian man­i­folds,” Ann. Math. (2) 47 : 1 (January 1946), pp. 85–​121. MR 0015793 Zbl 0060.​41416

[11] article S.-S. Chern: “Some new view­points in dif­fer­en­tial geo­metry in the large,” Bull. Am. Math. Soc. 52 : 1 (1946), pp. 1–​30. MR 0021706 Zbl 0063.​00834

[12] article S.-S. Chern and Y.-F. Sun: “The im­bed­ding the­or­em for fibre bundles,” Trans. Am. Math. Soc. 67 (1949), pp. 286–​303. MR 0032996 Zbl 0037.​10203

[13] article S.-S. Chern: “On the kin­emat­ic for­mula in the Eu­c­lidean space of \( n \) di­men­sions,” Am. J. Math. 74 (1952), pp. 227–​236. See also Boll. Uni­one Mat. It­al. 2 (1940). MR 0047353 Zbl 0046.​16101

[14] article S.-S. Chern and R. K. Lashof: “On the total curvature of im­mersed man­i­folds,” Am. J. Math. 79 : 2 (April 1957), pp. 306–​318. MR 0084811 Zbl 0078.​13901

[15] article S.-S. Chern and R. K. Lashof: “On the total curvature of im­mersed man­i­folds, II,” Mich. Math. J. 5 : 1 (1958), pp. 5–​12. MR 0097834 Zbl 0095.​35803

[16] incollection S.-S. Chern: “Min­im­al sur­faces in an Eu­c­lidean space of \( N \) di­men­sions,” pp. 187–​198 in Dif­fer­en­tial and com­bin­at­or­i­al to­po­logy (A sym­posi­um in hon­or of Mar­ston Morse). Edi­ted by S. S. Cairns. Prin­ceton Math­em­at­ic­al Series 27. Prin­ceton Uni­versity Press, 1965. MR 0180926 Zbl 0136.​16701

[17] article S.-S. Chern: “On the kin­emat­ic for­mula in in­teg­ral geo­metry,” J. Math. Mech. 16 (1967), pp. 101–​118. MR 0198406 Zbl 0142.​20704

[18] article S.-S. Chern and R. Os­ser­man: “Com­plete min­im­al sur­faces in Eu­c­lidean \( n \)-space,” J. Anal. Math. 19 : 1 (1967), pp. 15–​34. MR 0226514 Zbl 0172.​22802

[19]R. Bott and S. S. Chern: “Some for­mu­las re­lated to com­plex trans­gres­sion,” pp. 48–​57 in Es­says on to­po­logy and re­lated top­ics: Mé­m­oires dédiés à Georges de Rham [Es­says on to­po­logy and re­lated top­ics: Mem­oirs ded­ic­ated to Georges de Rham]. Edi­ted by A. Hae­fli­ger and R. Narasim­han. Spring­er (New York), 1970. MR 0264715 Zbl 0203.​54202 incollection

[20] incollection S.-S. Chern: “On the min­im­al im­mer­sions of the two-sphere in a space of con­stant curvature,” pp. 27–​40 in Prob­lems in ana­lys­is: A sym­posi­um in hon­or of Sa­lomon Boch­ner (Prin­ceton, NJ, 1–3 April 1969). Edi­ted by R. C. Gun­ning. Prin­ceton Math­em­at­ic­al Series 31. Prin­ceton Uni­versity Press, 1970. MR 0362151 Zbl 0217.​47601

[21] S.-S. Chern and J. Si­mons: “Char­ac­ter­ist­ic forms and geo­met­ric in­vari­ants,” Ann. Math. (2) 99 : 1 (January 1974), pp. 48–​69. MR 0353327 Zbl 0283.​53036 article

[22] article S.-S. Chern and J. K. Moser: “Real hy­per­sur­faces in com­plex man­i­folds,” Acta Math. 133 : 1 (1974), pp. 219–​271. MR 0425155 Zbl 0302.​32015

[23] article S.-S. Chern and S. I. Gold­berg: “On the volume de­creas­ing prop­erty of a class of real har­mon­ic map­pings,” Am. J. Math. 97 : 1 (1975), pp. 133–​147. MR 0367860 Zbl 0303.​53049

[24] book S.-S. Chern: Se­lec­ted pa­pers, vol. 1. Spring­er (New York), 1978. With a fore­word by Hung Hsi Wu and in­tro­duct­ory art­icles by An­dré Weil and Phil­lip A. Grif­fiths. MR 514211 Zbl 0403.​01012

[25] incollection S.-S. Chern: “Mov­ing frames,” pp. 67–​77 in The math­em­at­ic­al her­it­age of Élie Cartan (Ly­ons, France, 25–29 June 1984). As­térisque. 1985. Numéro Hors Série. MR 837194 Zbl 0614.​53038

[26] article T. E. Cecil and S.-S. Chern: “Taut­ness and Lie sphere geo­metry,” Math. Ann. 278 : 1–​4 (1987), pp. 381–​399. Ded­ic­ated to Friedrich Hirzebruch on the oc­ca­sion of his six­tieth birth­day. MR 909233 Zbl 0635.​53029

[27] book S.-S. Chern: Se­lec­ted es­says by S. S. Chern. Sci­ence Press (Beijing), 1989. In Chinese.

[28] book S. S. Chern: Se­lec­ted pa­pers, vol. II. Spring­er (New York), 1989. MR 1007138 Zbl 0682.​01017

[29] book S. S. Chern: Se­lec­ted pa­pers, vol. IV. Spring­er (New York), 1989. MR 1019829

[30] book S. S. Chern: Se­lec­ted pa­pers, vol. III. Spring­er (New York), 1989. MR 1025503