You are on page 1of 23

EARTHQUAKE ENGINEERING:

DESIGNING UNSEEN TECHNOLOGY AGAINST INVISIBLE


FORCES

James C. Williams

On 18 April 1906, the earth's crust dramatically shifted at the San


Andreas Fault, running north-south along the coastal region of
northern California. In San Francisco invisible shock waves threw
timber houses off their foundations, tumbled stone and brick
buildings, swayed bridges and towers, folded streets, uplifted
railway roadbeds, and collapsed tunnels. Gas from unseen yet torn
pipes ignited, fire swept the stricken city, and the fractured mains of
an invisible water system could not deliver sufficient water to douse
the flames. Some three thousand people died - crushed beneath
structures they had thought sturdy, struck by falling debris as they
fled into the streets, consumed by flames before they could dig or be
dug out from under rubble. Disturbing aspects of urban life that
normally remained 'out-of-sight, out-of-mind' suddenly became
visible when falling walls exposed busy brothels and overcrowded
tenements as well as flimsily constructed and inadequately
fireproofed buildings. Although people had experienced serious
earthquakes before, never had one so devastated a modern urban
centre. As Richard L. Humphrey, investigator for the United States
Geological Survey, put it: The whole civilised world stood aghast at
the appalling destruction which visited the city of San Francisco and
vicinity'.^
This essay addresses an aspect of invisible technology. Earthquake
engineering, a field born of the San Francisco cataclysm, provides
clues as to what might be considered salient features of at least one
form of invisible technology. First, it is a field which focuses on an
engineering response to natural forces which are themselves unseen
and therefore extremely difficult to measure or predict. Second, since
the engineering response is dependent on knowledge of seismic
forces for which hard data and theoretical understanding often have
been subjective and always incomplete, earthquake engineers have
faced sceptical opposition and been an almost indiscernible fringe
field of the engineering profession. Third, the engineering response
to earthquakes is in the form of aseismic building design, technology
James C. Williams 173

buried within structures. Because it is unseen and little understood


by the general public and because government cannot oversee every
building project, builders and property owners can skimp on
aseismic design and construction and one may never know the
difference. Its invisibility diminishes social consciousness. The
success of earthquake engineering only becomes evident when
seismic forces reveal both that it had been used and that it worked.
Earthquakes have interested human beings for centuries, and
anecdotal accounts of seismic activity appear in numerous historical
records. In 132 A.D., Chinese philosopher Chang Heng made the
earliest known earthquake recording instrument. Over a millennium
later, European natural philosophers and instrument makers were
drawn to the phenomenon. Robert Hooke lectured on earthquakes
and geology before the Royal Society in 1667-1668, and during the
next century a number of individuals designed devices to measure
tremors. In Italy, in 1783, the Calabrian earthquakes, a series of six
severe shocks, prompted the first official seismic investigation
commission. The nineteenth century saw further development in
seismological instrumentation as well as increased interest in
earthquakes by engineers. English civil engineer Robert Mallet,
called the 'primeval earthquake engineer' by California Institute of
Technology engineering professor George W. Housner, coined the
words 'seismology' and 'epicentre' plus other earthquake related
terminology. He was the first person to explode dynamite so as to
establish approximate earthquake wave velocities by measuring the
speed of elastic waves in surface rocks. He analysed and wrote about
damage wrought by a destructive 1857 earthquake in Naples and
compiled the first extensive catalogue of earthquakes.^
After 1875, developments in seismology occurred in Japan,
primarily through the efforts of British engineers teaching at Tokyo's
Imperial College of Engineering. There mining engineer John Milne
and mechanical engineers James Ewing and Thomas Gray became
interested in earthquakes. Ewing and Gray designed horizontal and
vertical seismographs that greatly advanced earthquake recording
and, with Milne, conducted elastic wave propagation experiments
using dynamite. Although Ewing and Gray returned to England (in
1883 and 1881 respectively), Milne stayed on for several more years.
He conducted experiments that measured seismic waves on both
hard and soft ground and compared motion on the floors of two-
story brick and wood buildings. He also helped found the
Seismological Society of Japan in 1880. With more foreign than
Japanese members and its transactions published in English, interest
174 Earthquake Engineering

in the society could not be sustained in its homeland. It lasted for


only twelve years, although Milne kept editing its 'transactions' as
the Seismological Journal of Japan until 1895, when he returned to
England to establish a seismological observatory on the Isle of
Wight.'
During this same period, earthquake study emerged in North
America. In 1868, a strong tremor on California's Hayward Fault
running through the east bay towns of Berkeley and Oakland caused
considerable damage in the San Francisco Bay area. A number of
brick buildings erected on landfill in San Francisco's commercial
centre were wrecked or badly damaged and thirty people died.
Interest in seismology surfaced among geologists associated with the
California Geological Survey, U.S. Coast Survey, and U.S. Geological
Survey. George Davidson of the Coast Survey served on a committee
investigating the Hayward earthquake, and after 1870 other scientists
published papers describing faults and tremors in California as well
as other North American locations. The first American seismographs
were installed in 1887 at the University of California campus in
Berkeley and at Lick Observatory, on Mount Hamilton some fifty
miles to the south. By 1900, there were several other North American
seismographic stations.^
The Hayward earthquake also damaged sufficient property, killed
enough people, and drove down real estate values sharply enough to
stimulate a few bay area property owners and contractors to build
stronger foundations and use structural reinforcement in new build-
ings. For example, San Francisco's leading financier William Ralston
reinforced the brick walls of his Palace Hotel, built in 1875, with some
three thousand tons of 'earth-quake proof iron banding. Masonry
cross walls designed into the hotel plus use of exceptionally good
mortar also gave the structure superior stability. A decade later, San
Francisco concrete block manufacturer Ernest L. Ransome, in
collaboration with architect George W. Percy, began to build
structures with reinforced concrete, a new and experimental
material. In 1889, Leland Stanford commissioned them to use it in
constructing a girls' dormitory and a museum at his new university
just south of San Francisco. After the 1906 calamity, these and other
buildings erected with earthquakes in mind became the first
structures from which engineers and architects could evaluate
aseismic building techniques.^
In the rural and agrarian world that characterised human societies
through most of the nineteenth century, peril from earthquakes
seldom concerned people. In the few cities that existed capital
James C. Williams 175

investments were relatively small, infrastructures simple, and people


independent of centralised services. Residents often got their own
water and disposed of their own wastes. They weathered storms in
their own cellars. When fire - the most common danger - broke out,
volunteers extinguished it. Since few disasters caused tragic
irreversible losses, not surprisingly only the most well-to-do bay area
investors undertook added building costs or experimented with
untested aseismic technology after the Hayward tremor. Most San
Franciscans continued disclaiming danger from earthquakes, fearful
their own economic fortunes would decline if word got out that the
forecast for San Francisco was, as Mark Twain put it, 'Shaky'.
'Occasional shakes, followed by light showers of bricks and
plastering'.^ Thirty years later. University of California professor
Andrew C. Lawson alleged that following the 1868 tremor the report
of 'a committee of scientific men' assembled by a prominent San
Francisco businessman 'was suppressed by the authorities, thru the
fear that its publication would damage the reputation of the city'.^
By the outset of the twentieth century, the nature of cities had
changed. Capital investments in modern cities were huge,
infrastructures complex, and citizens increasingly dependent on
centralised services for water, fuel, food, transportation,
communication, and even shelter. The nature of the modern city
enormously heightened the risk of ruinous loss to human life and
property from natural disasters. The powerful and unpredicted
forces that wreaked havoc on San Francisco in 1906 illustrated this
and provoked sustained engineering and scientific interest in
earthquakes.^ Gradually a culture of prevention evolved. Engineers
and geologists nurtured it at first, until those who invested in the
modern city - capitalists and governing officials - realised how
essential understanding earthquakes was to their pocketbooks.
Eventually they became sustainers of the culture of prevention,
drawing their needed expertise from the small cadre of earthquake
engineers who focused on the design of structural fortifications that,
once incorporated, remained invisible within the buildings and
structures they safeguarded.
Within days after the April 1906 earthquake, committees of
scientific men - geologists as well as engineers - formed to study its
effects. Engineers comprising the San Francisco Association of
members of the American Society of Civil Engineers (ASCE)
organised themselves to investigate seismic impact on buildings and
other constructions. Several of them also served on the city's
Committee of Forty on Reconstruction, while other individual en-
176 Earthquake Engineering

gineers participated in special studies by the State of California and


U.S. Geological Survey. In May, those among them who specialised
in building construction - they were just beginning to call
themselves 'structural engineers' - founded the Structural
Association of San Francisco, which evolved eventually into the
Structural Engineers Association of Northern California. In August, a
few engineers also joined geologists and other scientists in founding
the Seismological Society of America.^
Geologists concluded that building damage 'depended chiefly
upon the geological character of the ground'. Structures on hard
ground withstood seismic forces better than those on landfill and
alluvial soil, and the scientists urged sustained and serious study of
the subject.'® Engineers went further. 'A bridge or viaduct is planned
in the certainty that trains will run over it almost immediately it is
finished', said former California State Engineer William Hamilton
Hall. Therefore, 'a building for this Coast should be planned and
built with an equal sense that an earthquake may run under it quite
as promptly'. The solution rested not only in understanding the
properties of the ground on which one built, but in efficiently
designing and constructing buildings. The earthquake provided a
laboratory in which to study these problems. 'For the first time', said
Japan's foremost seismologist and earthquake engineer Fusakichi
Omori, engineers were able 'to study the effects of [seismic] shocks
on steel [framed] brick and reinforced concrete buildings' as well as
on 'ordinary brick, stone, wooden houses, bridges, water-pipes,
etc.'."
Engineers concluded that 'the earthquake ... demonstrated the
fact that San Francisco would have been safer had there been more
engineering design, supervision and [honest] building practice in the
city's making'.'^ From their studies of the city's remains, they con-
ceived basic guidelines for aseismic building design. Stone and
unreinforced masonry construction were completely unreliable.
Steel reinforced brickwork stood up best when bonded with concrete
mortar, not lime. Foundations, walls, and roof trusses needed to be
well-framed. Wood buildings did well if they were firmly secured to
solid foundations, used transverse framing, and possessed framing
continuity between floors. Tall buildings on soft ground seemed safe
if they had deep pile foundations. Reinforced concrete buildings
resisted seismic forces extremely well, and steel framed structures
withstood them even better, particularly with diagonal framing
recommended for wind stress. The ASCE local connmittee concluded
that any building designed for thirty pounds per square foot of wind
James C. Williams 177

pressure would have resisted the earthquake stress and persuaded


the city to include such a design requirement in its new building
ordinance. 'The prime requisite of the structure is elasticity', they
said, 'the ability of a structure to return to its original form after
distortion'.''
Engineers soon discovered that not everyone endorsed in their
recommendations. Since little was understood about seismic forces,
under the best of circumstances it was ticklish to draw conclusions
about earthquake-resistant design and construction from wrecked
buildings. In San Francisco, the proof itself was doubly muddled.
Following the earthquake, the great fire and efforts to control it by
dynamiting unwrecked buildings had destroyed or obscured much
of the seismic damage. Business rivals of engineers - architects,
brickmakers, and builders - were unwilling to give over the field to
the engineers or give up their own past building techniques based on
what they saw as unproved deductions drawn from subjective
investigation of dubious evidence.'^
Builders and brickmakers attacked the implication that past
building practices and materials were at the root of the calamity.
They claimed buildings withstood the earthquake if they had been
honestly constructed, with good materials, and all component parts
thoroughly tied to each other. Even though most brick buildings had
collapsed, the Palace Hotel remained as a monument to anyone
seeking evidence that building well was the key to earthquake-
resistant construction. Although the fire had destroyed its wood
interior, the hoters eight-storey-high brick walls survived.
Brickmakers attributed survival of the walls to superior materials and
workmanship. Engineers attributed their stability to design: iron re-
inforcing plus iron anchors that tied them to 'a honeycomb of brick
[cross] walls giving lateral stiffness in all directions' as well as to the
wood frame floors which acted as a diaphragm during the tremor
itself. The ASCE committee concluded that 'buildings of brick walls
and wooden interiors cannot be built which will not be wrecked in a
severe shock, it being a fault of design and not of materials or
workmanship'.'®
Brickmakers also lambasted engineers for promoting reinforced
concrete. Ernest Ransome, John B. Leonard, and a handful of other
engineers and architects had pioneered its use in California
construction and by 1905 were beginning to win over some
traditionalists. But brickmakers, joined by the Bricklayer's Union,
had held them off, lobbying the city not to permit reinforced concrete
construction in its building ordinances. From his position as an
178 Earthquake Engineering

associate editor of The Architect and Engineer of California, John


Leonard fielded editorials and articles which supported his
proponency of reinforced concrete and chided conservative officials
who were sceptical of the medium.^'^ But it was the earthquake that
won recognition for reinforced concrete, and along with a wind
pressure requirement, the city's new building ordinance approved
July 5, 1906, included approval for use of reinforced concrete plus a
requirement that brick construction be supported by steel framing. In
the year following the earthquake, seventy-eight reinforced concrete
buildings from one to ten-stories in height were started in the
burned-over part of San Francisco.^^
Like brickmakers and builders, many architects were put off by
criticism of past building practices by engineers, but more
importantly they resented engineers reproaching them for being
hide-bound and intruding into the field they considered their own.
But engineers did not yield. The architect', asserted William
Hamilton Hall, 'stood for all that was conservative and bound by
ancient precedent in building'. Engineers provided the technical and
practical innovations fundamental to modern high-class building
practice. Hence, they felt justified in requesting architects forego
outside building adornments and other embellishments that might
dislodge and fall during an earthquake and in asking them to put
assets into securely anchored cornices and top walls, into steel
framing for towers and steeples, and into structural fortifications
against high wind force rather than into decorations. Indeed, said
University of California engineering professor Charles Derleth, Jr.,
'engineers ... should insist upon their structural requirements ..., and
not allow the architect to injure the strength of the building for the
purpose of securing some less necessary architectural feature or
embellishment'.^^
Despite an atmosphere which might poison partnerships, some
engineers and architects, such as John Debo Galloway and John
Galen Howard, found it possible to work together and use new
media such as reinforced concrete. Before the tremor Galloway, a
civil engineer, had designed and constructed as a consultant to
architects the foundations, steel frames, and reinforced concrete
floors of three Class-A buildings in San Francisco: the Mutual Savings
Bank and the Shroth building - both ten-stories - and the twelve-
story Shreve building. He also contributed to design and
construction of the St. Francis Hotel. All survived the tremor, and he
found himself in demand as a structural engineer. In May 1906 he
formed a business partnership with Howard, supervising architect
James C. Williams 179

for the University of California. For almost three years, they designed
and constructed multi-storied buildings in San Francisco and Ber-
keley as well as the university's library building and Boalt Hall of
Law. Their winning alliance may well have stemmed from Howard's
previous experience in a successful partnership with New York civil
engineer Samuel M. Caldwell plus Galloway's earlier work for
architects. Additionally, Galloway's ego never was wrapped up in
earthquake engineering; he always considered hydroelectric power
and irrigation development his most important work.^*^
In the end, the property owners and investors continued to go to
architects for building design and construction. From the general
public's perspective engineers had no particular role to play in the
business. When people praised a building's appearance, the inevi-
table and obvious question was: 'Who was the architect?' William
Hamilton Hall regretted this but understood it:
'The architect's part of the result is visible - the style,
proportions, details and groupings, the selection of materials
and adaptation of parts, the special features and general ar-
rangement to the intended use. These are open to view and to
commendation or to criticism after the structure is finished. The
engineer's part is invisible. Hid below the ground or encased in
the shell and finish are the structural members - the bones,
cartilages and muscles, so to speak - of the great thing we call a
first class building.'
Of course, he added, 'If these are not all right [the building] may
tumble down some day'. This reality sparked the sense of social
responsibility that, despite the invisibility of their work, motivated
those engineers choosing to nurture the field of earthquake
engineering.^"
In California, earthquake engineering gains resulting from the
San Francisco tremor evaporated almost as quickly as they were
gotten. Even before rebuilding began, the powerful Southern Pacific
Company undertook a campaign denying that the earthquake had
caused the city's disaster. Panic selling on the New York Stock
Exchange drove down the value of the railroad's stock, significantly
endangering its borrowing power. With massive bonded
indebtedness, spending huge amounts of money at that moment to
contain the worst flood ever of the Colorado River, and confronted
with formidable earthquake damage repairs, the Southern Pacific,
according to San Francisco earthquake historians Gladys Hansen and
Emmet Condon, could not afford 'long-term losses of revenue that
180 Earthquake Engineering

might be fuelled by fear of future major West Coast earthquakes'.


Consequently, the company undertook a major public relations
campaign. 'Essentially', say Hansen and Condon, 'the Southern
Pacific Company began to rewrite the entire history of the disaster -
a simple and sanitised version - to diminish the impact of the
earthquake, and to assure easterners that investment in California
enterprises would continue to be good business'.^^
'The real calamity in San Francisco was undoubtedly the fire',
contended Southern Pacific's General Passenger Agent James
Horsburgh, Jr He led the railroad's public relations effort, filling the
firm's sleek Sunset Magazine from 1906 through 1908 with disinfor-
mational articles and illustrations by some of America's best writers
and photographers. He wrote chambers of commerce throughout
the state requesting they get lecturers and newspapers to downplay
the tremor 'Get them to make the story complete', said Horsburgh.
Have them show how little an area the earthquake really effected,
how quickly the city and state recovered, that the burned city was an
old city, and that only the most ancient part of the water-supply
system failed. Moreover, he said, assure people that serious seismic
activity is 'so infrequent that no city in the temperate zone has ever
been twice affected in a serious way by an earthquake'.^^
The Southern Pacific's version of the 1906 disaster soon became
the accepted one, repeated everywhere by newspapers and authors
of magazine articles and books, adopted by individuals,
organisations, and businesses. As people accepted that it was the fire,
not the earthquake, they began avoiding mention of anything
suggestive of earthquakes. It was as though government, business,
and citizens alike had embraced an unstated policy of concealment.
'It is feared that if the ground of California has a reputation for
instability', wrote G. K. Gilbert in Science magazine in 1909, 'the flow
of innmigration will be checked, capital will go elsewhere, and
business activity will be impaired'. John C. Branner, geologist and
Stanford University president as well as president of the
Seismological Society of America, observed in 1913 that 'this theory
has led to the deliberate suppression of news about earthquakes, and
even of the simple mention of them'.^'
Suppressing seismic danger in California seriously undermined
the progress of earthquake engineering. Many architects,
brickmakers, and builders simply ignored the conclusions reached
about aseismic construction by earthquake engineers. In late 1907,
amidst San Francisco's reconstruction, well-known architect Willis
Polk warned that 'many buildings are now being constructed in a
James C. Williams 181

manner that will court certain destruction in case of another


earthquake'. Other observers agreed and further accused property
owners and investors of seeking to cheapen construction so as to
secure greater return on their investments. City officials not only
averted their eyes from building ordinance violations, but in 1909
they acquiesced in pressure to weaken the thirty pound wind force
requirement developed by engineers in 1906, reducing it to fifteen
pounds. In 1910, they raised it back to twenty pounds but sixteen
years later again reduced it to fifteen pounds. Meanwhile,
earthquake engineers worried that too many structures 'were built
on the go-as-you-please basis by contractors who perhaps did not
know what really constitutes good building', and they called for a
state building code and inspection to supplement local ordinances.^^
Suppression slowed earthquake engineering progress even more
by making it seem that government support of seismic research
would unnecessarily squander pubUc funds. Neither the California
state nor the federal governments provided any funding for
university or independent research. Efforts failed to place a bureau of
seismology under the Smithsonian Institution or operate
seismographs through existing U.S. Weather Bureau stations. When
the Seismological Society of America foundered during its first ten
years, John Branner revived it through personal leadership and
procurement of an indispensable $5,000 gift. Stephen Taber, South
Carolina's state geologist, while a visiting lecturer on seismology at
Stanford University in 1920, bemoaned the situation: 'We have had
no government aid, and no state aid. The politicians want nothing to
do with it. Like the rest of the people, they try to forget it, adopting
the ostrich policy of burying their heads in the sand.'^
Thus seismologists - geologists and earthquake engineers alike -
found themselves in a difficult predicament. Both their
understanding of earthquakes and their development of aseismic
technology stemmed from very limited study of past earthquake
effects. Not only was their lack of solid information and
understanding about seismic forces and effects insufficient to
convince building professionals and politicians that their
recommendations should be followed, confidence in their expertise
was enfeebled even by fellow engineers. None other than perhaps
the best known engineer in California, William Mulholland, chief of
the Los Angeles Department of Water and Power, said in the wake of
a damaging tremor in Inglewood near Los Angeles in 1920 that, 'if an
earthquake comes, there is nobody on earth that can build an
aqueduct, or a building, that will be proof against it. You can build it
182 Earthquake Engineering

so that the damage will be slight; but it is going to rupture'.^'


Speaking before San Francisco's Commonwealth Club in 1925,
earthquake engineer Fienry Dewell suggested that 'with respect to
safe design against earthquakes, we have [not] only not progressed
but we have actually, seriously retrogressed'. The situation already
was changing, however, as new tremors reawakened Californians to
the importance of seismology and earthquake engineering. In 1923
the Kanto earthquake, estimated subsequently at 8.3 in magnitude
on the Richter Scale, shattered Tokyo and Yokohama. Approximately
100,000 people perished. Two years later a less serious but
nevertheless destructive tremor hit Santa Barbara, California. Both
earthquakes revitaUsed earthquake engineering and furthered
evolution of the culture of prevention. Investigation of the Japanese
cataclysm refined and added to knowledge about aseismic design
and construction. The Santa Barbara tremor jarred California
government and business interests out of their state of denial.^^
Immediately after World War I, Tokyo experienced a building
boom. Hundreds of reinforced concrete structures went up;
American as well as Japanese engineers designed and constructed
sixteen large steel-framed buildings. Yet, despite design and
construction that considered earthquake forces, the 1923 tremor
made it clear that aseismic technology needed more work. Ten of
Tokyo's sixteen new steel-framed buildings, for example, suffered
severe damage. Therefore, like San Francisco had before it, Tokyo
became an earthquake research laboratory for several months - an
especially valuable one in that fire had not obscured earthquake
effects as fully as it had in San Francisco. The ASCE formed a special
committee chaired by John Galloway to investigate it, and a number
of American engineers visited Japan and published their impressions
and conclusions. At California Institute of Technology, engineering
professor R. R. Martel began the first experimental seismic design
research in the United States, using a shaking table to correlate with
theoretical analyses the reaction of basic structural forms to
vibrations.^®
Tachu Naito, structural engineering professor at Waseda
University, Tokyo, later would recall that 'research and investigations
pertaining to earthquake resistant construction received added
impetus after the Kanto earthquake of 1923, including methods of
analysing building frames, computation methods for lateral forces,
theoretical study of building vibrations, the strength and optimum
location of bracing walls, characteristics of ground motion, [and]
dynamic or vibrational properties of soils ...'. According to Henry
James C. Williams 183

Dewell, Naito's own design work and his analyses of damage in the
Tokyo earthquake 'had a profound influence on structural design in
California and elsewhere'. He was the first to design buildings so that
horizontal force was distributed among structural elements in pro-
portion to wall areas or column spacing. He initiated practical
determination of relative wall, partition, and structural frame
rigidities. He illuminated 'the importance of a symmetrical
arrangement of resisting units'. All of these became fundamental
principles in aseismic design.^'^
The 1925 tremor that jolted California's coastal city of Santa
Barbara further stimulated efforts in seismic research revitalised by
the Kanto earthquake. Stanford University, for example, improved its
research program by installing a large shaking table. More important,
however, the Santa Barbara tremor aroused support for earthquake
engineering within the state's business community and among
goverrmient officials. Unlike the aftermath of the San Francisco
earthquake, there was no denial of disaster. In fact the Southern
Pacific Company, leader in the campaign to cover-up the impact of
the 1906 tremor, openly admitted the damage it suffered in 1925 and
praised the 'resourcefulness of the men upon whom fell the burden
of keeping the railroad running'. In September, the Commonwealth
Club, which its drew members from the business and professional
community, adopted a resolution that recognised the inevitability of
recurring earthquakes and urged the city to again revise its building
code and strictly enforce it. Across California, government felt
pressure from business interests and citizens. Planners of great
bridges and timnels across San Francisco Bay approached the
Seismological Society of America to help them locate earthquake
faults. The self-interest of the business community no longer rested
in concealing the danger of earthquakes; government no longer
could afford to not take action to safeguard their citizens against
future earthquake danger. Prevention of financial loss - and with it
loss of property and human life - became paramount. Business and
government leaders moved forward on two fronts, discovering in
each that they could make little progress without drawing on and
supporting the field of earthquake engineering.^
Prior to the 1920s, earthquake insurance was written almost
exclusively as a side line on fire insurance. In 1921, only $13,617 in
premiums for earthquake insurance per se were paid in the United
States; over 90 per cent of these policies were on CaUfornia
properties. In the year of the Kanto earthquake, however, premiums
paid on California properties jumped to $213,707. Although this
184 Earthquake Engineering

represented coverage of only a few million dollars, it symbolised the


changing business philosophy about earthquake risks. Bankers and
mortgagees began requiring earthquake insurance on large building
loans and corporate building-bond issues. The Santa Barbara tremor
accelerated this practice and caused a startling increase in the
demand for insurance. The net volume of earthquake insurance
written in California mushroomed to over $100 million in 1925 and
continued growing to $346 million by 1930. With a deductible of 10%
on total property value, it was 'drafted as to protect the banker or
mortgagee and the underwriter much more fully than it protect[ed]
the owner of the equity in the building'.^'
As investors sought to shelter their financial holdings through
insurance, underwriters learned how difficult it was to assess
potential earthquake hazard. Basic principles existed for earthquake-
resistant construction, but many aspects of aseismic design and
construction remained unsettled. Moreover, an incertitude
characterised scientific reports related to seismic activity. The
predictions of experts were fuzzy at best. In 1924, for example, the
U.S. Coast and Geodetic Survey mistakenly reported that southern
California hills had moved ten to twenty feet from positions in which
earlier surveys had placed them. This led some well-known
geologists to conclude that the pressures from such movement made
an earthquake very likely. The government survey blunder was
revealed in 1926, but not before apprehension among underwriters
had led to a sharp rise in earthquake insurance rates as well as to the
refusal by some companies to renew policies written just a year
before.^^
Nevertheless, since underwriters had somehow to assess
potential risk, they turned to earthquake engineers to help them as
much as possible. The Board of Fire Underwriters of the Pacific, for
example, retained Edward W. Bannister to help carry out a rating
survey in Los Angeles. Bannister supervised a force of inspectors
surveying 125 selected risks ranging from materials and methods of
construction to structural deficiencies. The successful initial survey
was extended to other cities along the California coast. By the end of
1926, the underwriters board had surveyed 2,721 structures and
began developing a 'means of fixing the measure of damageability by
definite calculations ... [and a way to] determine the possible
financial loss in an earthquake of specified severity From seismic
force and ground acceleration assumptions based on the San
Francisco, Tokyo, and Santa Barbara tremors, the board developed a
rating system to assess how well structures would withstand
James C. Williams 185

earthquake forces. In 1928, the underwriters board retained Tachu


Naito to provide details about the design, nature of damage, and
repair costs on modern structures in Tokyo effected by the 1923
Kanto tremor.^
Meanwhile, government began to take the peril of earthquakes
more seriously as rising insurance rates plus lender's earthquake
policy prerequisites on important new construction stirred business
organisations to seek better building ordinances. The San Francisco
peninsular city of Palo Alto was in a unique position to respond. The
city had grown up in symbiotic association with Stanford University.
When faced with issues concerning water, electric power, gas, or
other urban technological systems, Palo Altans always had drawn on
the expertise of Stanford professors, many of whom resided in the
community. Following the Santa Barbara tremor, the local chamber of
commerce turned to resident university experts, and found among
them leaders of the Seismological Society of America. Since its
founding the society had developed an international membership
and great influence in scientific aspects of seismology, but after the
Japanese calamity, its president, Stanford professor Bailey Willis, and
other California based members began steering it toward more
practical earthquake matters: 'building, construction, insurance,
investments, and human lives'. As they put it, 'The Seismological
Society of America is thus becoming more truly American than
heretofore'.^
The Society's Committee on Building for Safety had been
working since 1924 on a model general building code provision that
local governments could add to their building ordinances. It
contained lateral force requirements for design and construction of
buildings that represented the latest thinking of earthquake
engineers. Palo Alto's chamber of commerce took it to the city
council, which adopted it at the end of Summer 1925. Within a few
months Santa Barbara and a half-dozen other California
communities adopted the so-called 'Palo Alto Code', insurance
underwriters studied it as a basis for specific code regulations, and
the organisation of Pacific Coast Building Officials derived from it
provisions against earthquake stress in their own proposed U.S.
Pacific Coast Uniform Building Code, adopted at their sixth annual
conference in October 1927.^
In seeking to write building codes cognisant of earthquake risk,
government officials discovered, as had underwriters, that
information about earthquakes and aseismic building was
incomplete and often ambiguous. For example, Henry Dewell points
186 Earthquake Engineering

out that there were 'two schools of thought... [about] seismic design
of buildings. One held to the principle of rigidity combined with
strength as the essential factor; the other held that sufficient strength
and rigidity were difficult to attain, were unnecessarily expensive,
and that flexibility was not only an essential but also reduced the
amoimt of strength necessary to resist earthquake shock'. Those
favouring flexibility advocated making the first story of multi-storied
building 'so flexible in relation to the stiffness of the stories above
that the earthquake motion would be there absorbed'. It was not
discovered until the 1930s that a multi-storied building's first story is
usually more flexible than the upper stories anyway.^
Another issue that faced officials was how to effectively express
design requirements in building codes. Following the San Francisco
earthquake, engineers believed that a building designed for a heavy
wind shear also would resist earthquakes. Yet, as Japanese civil
engineer L. H. Nishkian noted in 1927, the two forces were
fundamentally different - one limited in magnitude and applied only
above ground, the other unlimited in magnitude and applied
underground. After the 1908 Messina earthquake, Italians developed
a quantitative seismic code that expressed lateral force acceleration as
a percentage of acceleration due to gravity — buildings had to
withstand a lateral acceleration of 8% g. Following the 1923 Kanto
earthquake the Japanese Home Office followed suit, adopting a
seismic coefficient of 10% g. In 1925 authors of the 'Palo Alto Code' as
well as designers of the Pacific Coast Building Officials proposed
code also embraced it. Agreement on exactly what the coefficient
should be and what variables should go into it, however, proved
more difficult, particularly since earthquake ground accelerations
had never been recorded. Although, engineers needed and wanted
such information, seismologists apparently either had not been
particularly interested in it or had been imsuccessful in designing
instruments to measure it.'^
In 1929, John R. Freeman, past president of the ASCE and
American Society of Mechanical Engineers as well as president of the
non-profit Manufacturers Mutual Fire Insurance Company, attended
the Tokyo World Engineering Congress. Now in his mid-seventies he
had, during the course of his career, developed an interest in
earthquakes. He had served on California's 1906 State Earthquake
Investigation Commission, worked on Pacific Coast building
projects, served on three commissions studying engineering issues
surrounding construction of the Panama Canal, and seen
earthquakes as an insurance problem. In Tokyo he met Cal Tech's R.
James C. Williams 187

R. Martel and Kyoji Suyehiro, Professor and Director of Tokyo


University's Earthquake Engineering Institute. 'He immediately
understood the need for a strong-motion accelerograph', recalled
George Housner, and on his return to the United States convinced
the federal government to undertake a program to design, construct,
and install the instruments. The Seismological Field Survey of the
U.S. Coast and Geodetic Survey got the assignment in 1931 and
placed the first accelerographs in southern California late the next
year.^
On 10 March 1933, accelerographs in Long Beach took the first
direct measurements of damaging earthquake ground motion. 'For
the first time', said Housner, 'engineers could see the nature of strong
ground shaking: the ampUtude of motion, the frequency
characteristics, and the duration of shaking'. In light of the striking
results, the 'strong motion program' was expanded and renamed the
California Seismological Program. 'Accelerographs, displacement
meters, and seismographs were installed at strategic points
throughout California and other Western states', reported Henry
Dewell, 'and periods of both natural and forced vibrations were
determined for buildings, bridge piers, water tanks, and dams'.
During the 1930s, hundreds of measurements were made, forced
vibrations created with 'a portable ground and building shaker'
developed by Stanford engineering professor Lydik S. Jacobsen.'^ At
last engineers had solid data from which to develop aseismic
building designs. It was a far cry from trying to develop aseismic
design based on motion descriptions such as that given by }.
MacGowan in the Transactions of the Seismological Society of Japan for an
1878 tremor in China: 'shaking was felt for the space of time taken in
swallowing 1/2 bowl of rice'.^
The Long Beach earthquake not only provided ground motion
data but impelled government and business to further support
seismological work and develop protective programs. Many modern
structural steel and reinforced concrete framed office buildings stood
up well in Long Beach, but school buildings suffered badly. Public
outcry prompted the state legislature to pass the Field Act a month
later, giving responsibility for overseeing building and rebuilding of
all school buildings in California to the State Division of Architecture
and mandating earthquake safety standards. The division adopted a
lateral force coefficient of ten per cent and other engineering
provisions from the 1933 edition of the Pacific Coast Building
Officials Conference Uniform Building Code, while the legislature
passed another bill, the Riley Act, requiring all other nonfarm
188 Earthquake Engineering

buildings in California be constructed to withstand a lateral force of


two per cent. Meanwhile, the Board of Fire Underwriters of the
Pacific hired earthquake engineers to create updated recommen-
dations for aseismic building design, which it published in 1935."^^
World War II diverted attention and resources everywhere from
seismology and earthquake engineering. In 1948, however, the San
Francisco Section of the ASCE and the Structural Engineers
Association of Northern California formed a joint committee to study
lateral forces. They reopened the question of determining
appropriate lateral force coefficients for building design. Chaired by
Standard Oil Company of California's chief civil engineer John
Rinne, who had been mentored during the early 1930s by Henry
Dewell and Walter Huber, the committee set about translating
acceleration data from the 1933 Long Beach earthquake and from the
1940 El Centro, California tremor into 'a rational dynamic method for
establishing design forces on a structure'. The committee tested their
calculations against the 1949 Los Angeles and 1948 San Francisco
building codes for multi-storied buildings, finding neither
completely satisfactory. Their report, first presented to the ASCE in
1951, recommended a minimum lateral force base coefficient of two
per cent to a maximum of six per cent and offered a model Lateral
Force Code.""
Cal Tech professors R. R. Martel, George W. Housner, and J. L.
Alford opposed the committee's conclusions. They took a
conservative approach to establishing lateral force coefficients and
felt the committee's had wrongly analysed the available data. The
insurance industry, with which Martel consulted, also opposed lower
lateral force coefficients, recommending coefficients from eight per
cent to 13 per cent and claiming that reducing them 'for tall buildings
to 4 per cent or less, as is allowed by some codes, is based on
economic expediency, rather than on positive theoretical
knowledge'. The committee chastised their opponents for academic
conservatism: 'The establishment of these coefficients is not an
academic problem but requires the judgement of those members of
the profession who are regularly engaged in the practice of structural
engineering'. The City of San Francisco, either agreeing with the
committee or bending to pressure from citizens seeking economic
expediency, soon adopted a new building code that embodied the
committee's method for determining coefficients.^^
On the fiftieth anniversary of the 1906 earthquake, a small group
of engineers assembled to hear and discuss some forty papers at the
First World Conference on Earthquake Engineering at the University
James C. Williams 189

of California in Berkeley. Like the forces with which these engineers


contended, their field remained almost invisible, a barely discernible
spot on the periphery of the engineering profession. Yet, times were
changing. The sort of theoretical work undertaken by the Joint
Committee on Lateral Forces had begun to move their field from 'an
experience-based study of past earthquake effects' toward 'a science-
based engineering discipline, with an organised body of knowledge,
a research program to add to that knowledge, and a working
interaction with the basic sciences of geophysics and seismology on
the one hand, and with the practising design and construction of
engineering works on the o t h e r M o r e o v e r , the culture of
prevention which from the start impelled support for seismology
and earthquake engineering in the larger community began growing
step-for-step with world-wide prosperity during the last half of the
twentieth century.
A quarter-of-a-century later, in 1984, the Eighth World Conference
on Earthquake Engineering returned to Berkeley. 'Instead of the
forty papers from eleven countries at the First Conference',
remembered Donald Hudson, 'there were now some eight hundred
papers from 54 countries'.'*^ Seismic forces, although still
unpredictable, no longer seemed quite so invisible. Earthquake
accelerations were seen more and more clearly with the aid of
advanced instrumentation. New and more reliable data gathered
with each new tremor permitted surer aseismic building design. The
work of earthquake engineers found its way into most new
structures in economically prosperous, earthquake prone parts of the
world, and investors in the modern city and people in general
expected buildings, structures, and infrastructures - especially newly
constructed office towers, apartment complexes, and highway
bridges - to be earthquake-proof.
In January 1994, a 6.8 magnitude earthquake struck Northridge,
California. Buildings and structures toppled on people, killing them
outright or leaving them to dig or be dug out from imder rubble. Gas
mains broke and ignited, water systems failed, supposedly seismical-
ly safe highway bridges collapsed, crowded living conditions of poor
immigrant families were revealed, and as one journalist reported, the
quake bared America's $3 billion X-rated video industry. After the
dust had settled, property owners and investors submitted $6.5 bil-
lion in insurance claims, far more than the $2.4 billion collected in
industry-wide premiums during the preceding five years. Insurance
companies and underwriters who had argued not so many years
before that 'the public has a right to real and not apparent safety
190 Earthquake Engineering

against earthquakes' began withdrawing from the earthquake


insurance business. And workers in Los Angeles who thought 'that
rollers under their office towers let them sway harmlessly when the
earth moves' became less smug when they discovered that 'a high-
rise on rollers' was just another urban myth.^
The culture of prevention girded by faith in technology had been
so successful that it led to a false sense of security. Even though
aspects of earthquake engineering and the seismic forces with which
its practitioners wrestled had in some ways become visible, a critical
feature of it as an invisible technology had not. The results of the
earthquake engineer's work - the technology buried within
buildings and structures to safeguard them against tremors -
remained unseen and rarely tested. In the end, even if one could be
sure aseismic design really was part of a structure, one could not be
sure it would work. For all its visibility, the essential attributes of
earthquake engineering continued to be invisible.

NOTES

1. U.S. Geological Survey Bulletin 324, The San Francisco Earthquake and Fire of April 18,
1906 and Their Effects on Structures and Structural Materials, Washington, D.C. 1907,
14. The best general account of the 1906 San Francisco earthquake is Gladys Hansen
and Emmet Condon, Denial of Disaster (San Francisco, 1989).
2. James Dewey and Perry Byerly, 'The Early History of Seismometry (to 1900)',
Bulletin of the Seismological Society of America, 1969, 59: 183-227; G. W. Housner,
Tiistorical View of Earthquake Engineering', Proceedings of the Eighth World
Conference on Earthquake Engineering, Post-Conference Volume (Englewood Cliffs, N.J.,
1986), 27-9. A good example of a late-nineteenth century book about earthquakes is
Arnold Boscotwitz, Earthquakes (trans. C. B. Pitman) (London, 1890).
3. Ibid., 29; Dewey and Byerly, op. cit. (2), 195-211. Professor Fusakichi Omori, who led
Japanese efforts in seismology, let the society's journal expire after Milne departed.
Henry D. Dewell, 'Progress of Earthquake-Resistant Design', Civil Engineering,
1939, 9:601-604, has divided development of earthquake engineering into four
periods: 1880-1906 (formation of the Japanese society to the San Francisco
earthquake), 1906-1923 (Tokyo earthquake) or 1925 (Santa Barbara earthquake), 1923
or 1925-1933 (Long Beach earthquake), and 1933-the present (publication of his
article in 1939). I generally concur with Dewell but push back the first period, end
the second with 1923, extend the fourth to 1956 (year of the first World Conference
on Earthquake Engineering), and suggest a contemporary period from that date.
For a sense of issues and resources in seismology, see Edward P Hollis, Bibliography
of Engineering Seismology (2nd ed., Berkeley, 1958).
4. California, State Earthquake Investigation Commission, The California Earthquake of
April 18, 1906 (Washington, D.C., 1908), l:vii-viii and 434-451; Michael L. Smith,
Pacific Visions: California Scientists and the Environment, 1850-1915 (New Haven, 1987),
64. Clarence Edward Dutton was perhaps the first American to write a general text
on earthquakes: Earthquakes in the Light of the New Seismology (London, 1904).
5. Hansen and Condon, op. cit. (1), 8-11 and 64; U.S. Geological Survey Bulletin, op. cit.
(1), 112-114 and 149-150; Joint Committee on Lateral Forces, San Francisco Section,
American Society of Civil Engineers, and Structural Engineers Association of North
James C. Williams 191

California, 'Lateral Forces of Earthquake and Wind', Transactions of the American


Society of Civil Engineers, 1952,117:720. On Ransome and reinforced concrete, see C.
W. Whitney, 'Ransome Construction in California', The Architect and Engineer of
California, 1908, 12:n.p. between 48-49; John W. Snyder, 'Buildings and Bridges for
the 20th Century', California History, 1984, 68:281-282; Carl W Condit, American
Building (Chicago, 1968), 171-172. Ransome discusses his own work in Ernest L.
Ransome and Alexis Saurbrey, Reinforced Concrete Buildings, (New York, 1912).
6. Quoted in Hansen and Condon, op. cit. (1), 8. Head of the California Geological
Survey, Josiah D. Whitney, observed in 1868 that Californians possessed an attitude
of 'assumed indifference' to earthquakes. See Harry O. Wood, 'The Earthquake
Problem in the Western United States', Bulletin of the Seismological Society of America,
1916, 6:199.
7. State Earthquake Investigation Commission, op. cit. (4), 434. During the 1980s,
California seismologists again became interested in this issue. W. H. Prescott,
'Circumstances Surrounding the Preparation and Suppression of a Report on the
1868 California Earthquake', Bulletin of the Seismological Society of America, 1982,
72:2389-2393, published a letter written in 1908 by U.S. Coast Survey scientist
George Davidson in which he claimed a report had been prepared but suppressed.
In 1986 Michele Aldrich, Bruce A. Bolt, Alan E. Levitón, and Peter U. Rodda,
published 'The 'Report' of the 1868 Haywards Earthquake', Bulletin of the
Seismological Society of America, 1986, 76: 71-76, in which they concluded 'that
probably no major or integrated report on the 1868 earthquake was ever issued ...
et alone suppressed'. Likely no report was prepared, but that in itself may be
evidence that a sort of repression occurred. There appear to have been conflicts and
disagreements between members of the 1868 committee, and these well may have
been related to interpreting earthquake danger and damage in terms of the city's
reputation.
8. Two years after San Francisco's estimated 8.3 magnitude earthquake, 83,000 people
died in a tragic 7.3 magnitude tremor in Messina, Italy. An impetus toward building
codes which established quantitative seismic construction standards resulted from
it, leading G. W Housner, op. cit. (2), 30, to suggest that the Messina, not the San
Francisco earthquake resulted in 'the birth of practical earthquake design of
structures'. Notwithstanding improved building code developments stemming
from the Messina tremor, the San Francisco earthquake kindled the field of
seismological engineering. As San Francisco structural engineer H. J. Brunnier
observed in 1956, it 'accelerated ... research in earthquake engineering and . . .
progress has been steady and in the direction of rational and reasonable design
criteria'. See Brunnier, 'The Development of Aseismic Construction in the United
States', Proceedings of the World Conference on Earthquake Engineering (Berkeley, 1956),
1:26-2. Also see William Herbert Hobbs, Earthquakes: An Introduction to Seismic
Geology (New York, 1907), who credits the 1906 tremor with 'a lively interest in the
subject' (iv).
9. Report of the General Committee and Six Special Committees, 'The Effects of the
San Francisco Earthquake of April 18, 1906, on Engineering Constructions',
Transactions of the American Society of Civil Engineers, 1907, 59: 208-329; State
Earthquake Investigation Commission, op. cit. (4); U.S. Geological Survey Bulletin,
op. cit. (1); 'Structural Association of San Francisco', The Architect and Engineer of
California, 1906, 5:49-50; Sidney D. Townley, 'John Casper Branner', Bulletin of the
Seismological Society of America, 1922,12:1-11. Carroll W Pursell, Jr. provides a general
sense of the engineering profession's development in San Francisco during this
period in 'The Technical Society of the Pacific Coast, 1884-1914', Technology and
Culture, 1976,17: 702-717.
10. State Earthquake Investigation Commission, op. cit. (4), 241.
11. Wm. H. Hall, 'Municipal Engineering-Its Relation to City Building', The Architect and
Engineer of California, 1906, 7:20; E Omori, 'Preliminary Note on the Cause of the
192 Earthquake Engineering

California Earthquake of 1906', in David Starr Jordan (ed.) The California Earthquake
of 1906 (San Francisco, 1907), 304. Hansen and Condon, op. cit. (1), 121, note that San
Francisco remained an active laboratory for earthquake study for at least three
years.
12. Hall, op. cit. (11), 22.
13. Report of the General Committee, op. cit. (9), 234 and 235-236. Charles Derleth, Jr.,
who served on the committee, elaborated on the committee's conclusions in
'Destructive Extent of the California Earthquake' in Jordan, op. cit. (11), 119-212. Also
see U.S. Geological Survey Bulletin, op. cit. (1), 71-76 and 142-145; and the May and
June 1906 issues of The Architect and Engineer of California for a number of articles in
which engineers and architects suggest these sorts of structural lessons from the
earthquake.
14. Hansen and Condon, op. cit. (1), 14-15 and 107-127.
15. Report of the General Committee, op. cit. (9), 225 and 236; Brunnier, op. cit. (8), 26-2;
'Brick Well Laid is All Right', The Architect and Engineer of California, 1906, 5:n.p.
16. Snyder, op. cit. (5), 283-284. Also see 'Brickmakers' Opposition to Re-enforced
Concrete', The Architect and Engineer of California, 1906, 5:68; Hansen and Condon,
op. cit. (1), 110-111.
17. 'The New Building Ordinances', The Architect and Engineer of California, 1906, 5: 70;
William Hamilton Hall, 'The Rebuilding of San Francisco-Reinforced Concrete
Buildings', The Architect and Engineer of California, 1907,9:61-67.
18. 'The New Building Ordinances', The Architect and Engineer of California, 1906, 5: 70;
William Hamilton Hall, 'The Rebuilding of San Francisco-Reinforced Concrete
Buildings', The Architect and Engineer of California, 1907, 9: 61-67.
19. John Debo Galloway papers, MS 71-826, No. 1, Water Resources Archive, University
of California, Berkeley; Report of the General Committee, op. cit. (9), 210; 'In
Memoriam: John D. Galloway', California Historical Society Quarterly, 1943, 22:187;
Loren W. Partridge, John Galen Howard and the Berkeley Campus: Beaux-Arts
Architecture in the Athens of the West' (Berkeley, 1978). Hall, op. cit. (17), 62, observed
that 'the architects most prominent before the public by reason of the notable
buildings with which their names have in the past been connected, are not
prominent as the architects of reinforced concrete buildings'.
The St. Francis Hotel and Shreve building stand today. During his partnership
with Howard, Galloway chaired the local ASCE members special committee on Fire
and Earthquake Damage to Buildings, was an active member of Committee of
Forty's subcommittee revising San Francisco's building laws, and served as mentor
to two young civil engineers-Walter L. Huber and Henry D. Dewell-who during the
1930s and 1940s came to rank among California's 'foremost earthquake engineers'.
(On Huber and Dewell, see Winfield Scott Downs (ed.). Who's Who in Engineering
(5th ed.. New York, 1941), 452 and 871, and Proceedings of the Fourth World Conference
on Earthquake Engineering (Santiago, Chile, 1969), 4: 49.)
In January 1907, architect Charles Hays and civil engineer A. H. Markwart joined
Howard and Galloway's firm. At the end of 1908, when the partnership dissolved.
Galloway started a partnership with Markwart, who, like he, was primarily
interested in hydroelectric power. Later, Galloway chaired the ASCE Special
Committee on Effects of Earthquakes on Engineering Structures formed following
the 1923 Tokyo earthquake and at age seventy-one published, with colleague
Leander M. Hoskins, 'Earthquakes and Structures', Transactions of the American
Society of Civil Engineers, 1940,105: 269-322.
20. Hall, op. cit. (11), 23. On engineering and social repsonsibility, see Edwin T. Layton,
The Revolt of the Engineers: Social Responsibility and the American Engineering Profession
(Cleveland, 1971).
21. Hansen and Condon, op. cit. (1), 108.
22. Ibid., 108-110.
23. Gilbert is quoted in Wood, op. cit. (6), 199; J. C. Branner, 'Earthquakes and Structural
James C. Williams 193

Engineering', Bulletin of the Seismological Society of America, 1913, 3: 2. Also see


Hansen and Condon, op. cit. (1), 110-111; Stephen Taber, 'Seismology in the United
States', Bulletin of the Seismological Society of America, 1911,1: 3.
24. Sumner Hunt, 'Building for Earthquake Resistence', Bulletin of the Seismological
Society of America, 1922,12:18; Willis Polk, 'The Reconstruction of San Francisco', The
Architect and Engineer of California, 1907, 10:37; O.A. Johnson, 'Says Owners Are at
Fault', The Architect and Engineer of California, 1906,5: n.p.; 'Owners Blamed for Poor
Construction', The Architect and Engineer of California, 1907,9:69; Brunnier, op. cit. (8),
26-3.
25. 'The Earthquake Problem in Southern California: Address by Stephen Taber',
Bulletin of the Seismological Society of America, 1920,10:288; Taber, op. cit. (23), 1-4; John
Ripley Freeman, Earthquake Damage and Earthquake Insurance (New York, 1932), 2;
Townley, op. cit. (9), 1.
26. 'The Earthquake Problem in Southern California: Address by William Mulholland',
Bulletin of the Seismological Society of America, 1920,10: 293.
27. Dewell is quoted in Hansen and Condon, op. cit. (1), 137-138. Earthquake magnitude
is expressed internationally by means of the the Richter scale, which was developed
by California Institute of Technology professor Charles Richter in 1935.
28. Dewell, op. cit. (3), 602-603. Among published reports on the Kanto earthquake, see
Tadashi Taniguti, 'Seismic Action and Damage in Relation to Character of Building',
Proceedings of the World Engineering Congress, Tokyo, 1929, 8:181-212; Joseph S. Ruble,
'The Earthquake in Japan', ibid., 210-220; and three articles by H. M. Hadley, district
engineer of the Portland Cement Association in Seattle, Washington: 'Earthquake-
proof Building Construction as Revealed by the Japanese Earthquake', Bulletin of the
Seismological Society of America, 1924,14: 6-8; 'The Japanese Earthquake and its Effect
upon Buildings', The Architect and Engineer of California, 1924, 77: 83-90; 'How
Structures Withstood the Japanese Earthquake and Fire', Proceedings of the Twentieth
Annual Convention of the American Concrete Institute, 1924, 20:188-209. Hollis, op. cit.
(3), 116, reports in his 1958 bibliography that the ASCE committee compiled a
comprehensive but unpublished report, the original of which is at the Engineering
Societies Library in New York City. I was unable to locate a microfilm copy
reportedly on file at the University of California Library
29. Dewell, op. cit. (3), 602; Tachu Naito, Tifty Years of Earthquake Engineering Practice',
Proceedings of the Second World Conference on Earthquake Engineering (Tokyo, 1960),
1:130. A brief biographical sketch of Naito is in Proceedings of the Fourth World
Conference, op. cit. (19), 4: 49.
30. W H. Kirkbride, 'The Earthquake at Santa Barbara, California, June 29, 1925, as It
Affected the Railroad of the Southern Pacific Company', Bulletin of the Seismological
Society of America, 1927,17:1; Henry D. Dewell, 'Earthquake-Resistant Design: Data
of Design', Engineering News Record, 1928,100: 651; Freeman, op. cit. (25), 693-694; 'To
All Who Are Interested in Earthquakes', Supplement to Bulletin of the Seismological
Society of America, 1925, 15: n.p. The first English language book on earthquake
engineering was pubUshed in New Zealand following the Santa Barbara
earthquake, and it drew particluarly on engineering investigations the San
Francisco, Kanto, and Santa Barbara tremors: C. Reginald Ford, Earthquakes and
Building Construction (Auckland, 1926).
31. Ibid., 14,12-13 and 15-22; 'To All Who Are Interested', op. cit. (30), n.p.; Dewell, op.
cit. (3), 603.
32. Freeman, op. cit. (25), viii, 19-22; Henry D. Dewell and Bailey Willis, 'Earthquake
Damage to Buildings', Bulletin of the Seismological Society of America, 1925,15: 282.
33. Edward W. Bannister, 'Rating Buildings for Liability to Earthquake Damage',
Engineering News-Record, 1927, 98: 1052 and 1053-1054; Henry D. Dewell, 'A Survey
of the 1923 Earthquake Damage to Buildings in Tokyo, Japan', Engineering News
Record, 1930,105: 51-5A.
34. 'To All Who Are Interested', op. cit. (30), n.p.; Dewell, op. cit. (3), 603; Wardell
194 Earthquake Engineering

Winslow, Mo Alto: A Centennial History (Palo Alto, California, 1993), 217-223.


35. To All Who Are Interested', op. cit. (30), n.p.; Freeman, op. cit. (25), 695-700. Before
Santa Barbara adopted the 'Palo Alto Code', community leaders had formed an
architectural advisory committee to oversee rebuilding the downtown and charged
that committee with amplifying its existing code concerning earthquake-resistant
construction. But old habits seemed to have lingered. Under the guise of earthquake
safety, the committee focused wholly on redesigning downtown Santa Barbara to
suit the architectural tastes of its members. See Bernhard Hoffmann, 'The Rebuilding
of Santa Barbara', Bulletin of the Seismological Society of America, 1925,15: 323-328.
36. Dewell, op. cit. (3), 602.
37. L. H. Nishkian, 'Design of Tall Buildings for Resistance to Earthquake Stresses', The
Architect and Engineer of California, 1927, 88: 73-83; Housner, op. cit. (2), 31; Freeman,
op. cit. (25), 696-699; Christopher Arnold, 'Seismic Design', in Joseph A. Wilkes and
Robert T. Packard (eds.). Encyclopedia of Architecture, Design, Engineering &
Construction, (New York, 1988-1990), 4: 401.
38. Housner, op. cit. (2), 32; Freeman, op. cit. (25), v-vi; Dewell, op. cit. (3), 604.
39. Ibid.-, Housner, op. cit. (2), 32; Donald E. Hudson, 'Nine Milestones on the Road to
Earthquake Safety', Proceedings of the Ninth World Conference on Earthquake
Engineering (Tokyo, 1989), 2:8. Hudson underscores that the first near-source
recording of a truly great earthquake (8.1 magnitude in the epicentral region) was
not until the Mexico City tremor of 1985. It is worth noting that by the mid-1950s
only seventy strong-motion accelerographs were installed world-wide. It was not
an inexpensive program.
40. Quoted in Housner, op. cit. (2), 33.
41. H. M. Engle and John E. Shield, Recommendations-Earthquake Resistant Design of
Buildings, Structures, and Tank Towers (rev. ed., San Francisco, 1950), 5. The forward
of the 1950 edition is the same as that in the 1935 edition. Also see Dewell, op. cit. (3),
603-604; Merle A. Ewing and Charles N. Herd, 'Criteria for Structural Design in
California Schools', Proceedings of the World Conference on Earthquake Engineering,
Berkeley 1956,1: 36-1 to 36-15.
42. Joint Committee on Lateral Forces, op. cit. (5), 716-754. A brief biographical sketch of
the committee chair John Rinne is in Proceedings of the Fourth World Conference, op. cit.
(19), 4: 49.
43. Joint Committee on Lateral Forces, op. cit. (5), 755-780; Brunnier, op. cit. (8), 26-5 to
26-6. The Pacific Fire Rating Bureau's position is in Engle and Shield, op. cit. (41), 18;
and the code provisions as adopted by San Francisco in 1956 are embodied in
Structural Engineers Association of Northern California, Lateral Force Building Code
Requirements (San Francisco, 1956).
44. Hudson, op. cit. (39), 2: 3.
45. Ibid., 2: 7.
46. The insurance industry quotation is from Engle and Shield, op. cit. (41), 19. On the
Northridge earthquake, see 'Killer Quake in L.A.', San Francisco Chronicle, 18 Jan.
1994,1; William Arnold, 'Quake damage to adult-video firms gives industry pause',
San Jose Mercury News, 25 Jan. 1994, ID; Steve Farr, 'Quake claims in crisis: Insurance
firms want U.S. plan', San Jose Mercury News, 19 July 1994, 3B; George de Lama,
'Shaky ground: Insurance crisis looms in wake of costly Northridge earthquake',
San Francisco Examiner, 24 July 1994, B4.
Jane E. Allen reported general belief in the myth of office towers on rollers in 'L.A.-
style urban legend is just that: No, there are no rollers under the high-rises', San Jose
Mercury News, 15 May 1994,3B. In fact a kernel of truth is behind the myth. During
the 1920s, earthquake engineer Henry D. Dewell, op. cit. (30), 653, reported on
Japanese experiments with free foundation design. 'Some small buildings and some
lighthouses in Japan', he said, 'have been provided with free foundations, by rollers
placed between foundations and superstructure... This construction has, however,
not yet been worked out in practical form for general application to large buildings'.

You might also like