A Brief History of Mining in Utah

Index For This Page

(Return to Mining Index page)

By

BURT B. BREWSTER, E. M.

Editor and Publisher, The Mining and Contracting Review

with a

FOREWORD By

D. C. JACKLING

Founder of the Utah Copper Enterprise, which revolutionized the low grade copper industry throughout the world-called by many "The Father of the Porphyry Coppers."

(From Utah, A Centennial History, Volume II, By Wain Sutton, editor, Lewis Publishing Co., New York, 1949; scanned and formatted for electronic publication by Don Strack, October 2003.)

Foreward

Mr. Burt B. Brewster's masterful review of mining history in Utah for one hundred years following the pioneers' initial settlement in the area since occupied by Salt Lake City has revived bright and happy memories of my participation in the progress of mining activities within the State throughout over half a century during which I have been continuously identified with mining and allied industries in that great commonwealth.

The search for and development of metallic mineral occurrences was retarded for the first few decades following the advent of the pioneers in the Salt Lake Valley for the compelling reasons that the population was small and the necessities of life meager, thus demanding almost sole devotion of settlement energies to the provision of shelter and the pursuit of agricultural production of sustenance for the people. and provender for domestic animals which were the only source of power available for tilling the soil and the conveyance of goods, either vegetable or mineral, awaiting the arrival of transportation means required for the shipment of even rich ores to points of conversion to metallic state.

I subscribe wholeheartedly to Mr. Brewster's expression of affectionate regard for, the State of Utah and high esteem for those responsible for the early development of its resources, including mineral kind. In this connection, I deem it relevant to emphasize the fact that, in successive eras, the inauguration and progress of mining as well as other of the State's industries are creditable in very large degree to resident citizens of Utah, a great proportion of whom were native-born descendants of the pioneers and their adherents.

Having thus briefed my hearty commendation of Mr. Brewster's chronicles, it is fitting that I should express gratification for the privilege of complying with his complimentary request that I prepare a foreword to accompany their publication, notwithstanding my misgivings of ability to contribute much of interest or substance to the value of his most impressive statement of facts and truthful analysis of events, constituting a chronological record on the subject of his writings such as never before has been assembled and could only be by one possessed of his inimitable talents, editorially, and ardent allegiance to the principles of truth, right and justice as his uncompromising integrity guides him to discriminate between reality and pretense.

D. C. JACKLING

San Francisco, California, May, 1948

Preface

As stated in the opening chapter of this necessarily condensed outline of mining during the one hundred years (1847-1947) of Utah history, only facts pertaining to production, the value thereof, with few references to individuals, are presented. Tabulations are included for purposes of comparison and statistical record.

Plate I, detailing the Utah Copper Enterprise, speaks for itself. It is published with consent of the copyright owner.

The roles of the railroads, power companies, and smelting companies in Utah mining cannot begin to be recounted in the space allotted. Their roles are left to other sections of this history of Utah.

To have enumerated and treated adequately the early and subsequent contributions of many men to Utah mining would require thousands of pages. Among them are such men as Daly, Dailey, Ellis, Sr. and Jr., Judge, Judd, Bamberger, Buchman, Fitch, Holden, Dickson, Allen, Keith, Kearns, Shapiro, Walker Brothers, Lamborne, Thompson, DeLamar, MacNeil, Penrose, Macvichie, Pett, Wall, Dern, Knight, Gemmell, Guggenheim, Newhouse, Hanchett, Hatch, Klepetko, Clark, Cates, Channing, Raddotz, Snyder, McIntyre, Moore, McChrystal, Elton, McCornick, Moffat, Shilling, MacDonald, the Hyland Brothers, Mudd, Krumb, Parsons (C. C.), Hammond, Hayden, Stone, George Bradley, the Janneys, Goodrich, Engelmann, Raymond, Tooker, Jennings, Cameron, Heiner, and many others not specifically mentioned in the text.

The reader must not mistake mere affection for the author's hope for a Foreword from Daniel C. Jackling. Cold-blooded contemplation of the importance of Utah's mining to the State (more than half the population depends upon it) and to the Nation (perhaps World War II could not have been won without it), its current and future role in history, left no other choice. Perusal of the context will confirm our choice.

Credit for much early factual data is due V. C. Heikes, who so concisely gave historical reviews of Utah districts in Professional Paper i i 1, The Ore Deposits o f Utah, United States Geological Survey, under the direction of Franklin K. Lane, Secretary of the Interior, and George Otis Smith, Director of the Survey.

Other credit is due T. A. Rickard's A History o f American Mining; A. B. Parsons in his The Porphyry Coppers; the United States Geological Survey; the United States Bureau of Mines; and The Mining and Contracting Review, the founder of which, Will C. Higgins, once wrote: "Were it not for mining, Salt Lake City would still be a whistle and water tank stop on a transcontinental railroad."

The westward expansion of industry, with its fabrication of metals, and other potential consumers of mineral products, promises a bright future for Utah provided the taxing agencies of the State are restrained from hamstringing industry and the venture capital which makes expansion possible. To live in Utah is to love Utah.

BURT B. BREWSTER

Salt Lake City, May, 1948

Introduction

The glamour, the romance, and the pioneer courage of early Utah mining would fill volumes. But, interesting as is the detail thereof, it cannot be treated adequately in the brief history here offered for the record. Nor is there an opportunity for technical discussion.

The major "eras" in Utah metal mining have been the high grade, direct smelter shipping days; the early concentration of complex ores; and the scientific attack upon the lower and very low grade ores, particularly those of copper.

In point of dollar value (see Table I), the nonferrous metals-copper, silver, lead, gold, and zinc-in the order named, have led all-time production in Utah. The production of quicksilver, arsenic, tungsten, manganese, antimony and bismuth has been of relatively minor importance.

Coal, for which no accurate value of production can be estimated, ranks next to the nonferrous metals in ultimate value.

Iron ore, the commercial production of which had its first appreciable impetus in 1924, is gaining rapidly in output because of the peacetime operation of wartime steel facilities in addition to an existing prewar pig iron plant and shipments of ore to California and Colorado steel mills.

Radium and uranium bearing vanadium ores hold a unique place in Utah mining, emphasized by the development of atomic energy, destructive and constructive.

Recovery of molybdenum from the copper ore of Bingham has been a comparatively recent accomplishment but has reached major proportions, placing the State second only to Colorado in world production of the metal.

Non metallics attract increasing attention.

Utah claims almost sole production of the hydrocarbons, gilsonite and ozokerite; and use of the State's natural rock asphalt has expanded during recent years.

Nonferrous Metals

As in the case of most Western states, nonferrous metal mining in Utah had its real inception in the recovery of gold and silver, the former from placer deposits, the latter from. ore on or near the surface. Silver-lead ores constituted the "bonanzas" of the early days, followed by copper. Zinc did not become profitable until metallurgical advance in its separation in treating complex ores made it valuable rather than a handicap to the miner. By-product gold from copper ore has made Utah a leading gold producing state.

Although the records state that silver-lead ore near Great Salt Lake was "known to the Mormons in 1857; that silver-lead ore was discovered at Gold Hill in 1859; and that a follower of Brigham Young melted lead ore in the i 85os," it appears that George Olgilvie's find of fragments of lead ore in or near Bingham Canyon in 1863 is what drew attention to the area and led to the forming of West Mountain (Bingham) Mining District in 1864, in which year placer gold was discovered at the mouth of Bingham Canyon.

Various historians credit the earliest Bingham Canyon activities to the influence of General P. E. Connor of the United States Army and his California volunteers, many of whom were prospectors. These same historians attribute Brigham Young's opposition to "Mormon" participation in mining to anxiety lest they desert their tilling of the soil with resultant loss of desperately needed food for the pioneers.

The districts most productive of nonferrous metals have been Bingham, Park City, Tintic, Ophir, Beaver County, Big and Little Cottonwood, Mercur, Washington County, American Fork, Piute County, Lucin, Fish Springs, Clifton, Park Valley and Promontory. But today the Bingham, Park City, Tintic and Stockton districts produce most of Utah's five principal metals.

Copper

Leader in total value of production and the inspiration for the most vital world strides in modern mining and metallurgy, copper deserves first consideration, despite the fact that it was overshadowed in importance until rgo4 by silver, lead, and gold. In fact, in the early days the red metal was considered a nuisance because of the treatment complications involved when it was associated with silver and gold. As a result much copper ore was considered to be "waste" and cast aside on mine dumps.

In order of copper output the Bingham District has led, by so wide a margin as to make all but the Tintic District's contribution appear insignificant. The rank of other copper producing districts has been Park City, Beaver County, Ophir, Lucin, St. George, and Big and Little Cottonwood, with scattering output from other districts such as the LaSalle area.

By far the greatest recovery of copper metal has been from copper ores.

The first shipment of copper ore in Utah is recorded as having been made from Bingham by Walker Brothers of Salt Lake City in 1868, but it was not until 1886 that the total annual value of copper output reached a quarter of a million dollars. In that year about 2,500,000 pounds, valued at about $267,000 was produced. From then until 1896, when 3,502,012 pounds valued at $378,217 were produced, the annual value of output rose and fell to a high of $358,oi6 in 1888, and to a low of $109,019 in 1894.

The acceleration in output of copper during the period 1890 to 1900 was due principally to the operation of the Highland Boy mine in Bingham and the Mammoth properties in the Tintic Districts.

Bingham Copper Production

Copper produced in the Bingham District from 1865 through 1947 totaled 5,166,536 short tons of metal.

Following the early initial large shipment of copper from the Highland Boy mine in 1896, the development of copper sulphide ore deposits increased and production rose to 6,196,660 pounds valued at $1,028,646 in 1900. In 1899 the first "modern" copper smelter in Salt Lake Valley had been built by the operators of the Highland Boy mine. Mines contributing to copper output at the turn of the century had been the Highland Boy, Boston Consolidated, Yampa, Old Jordan, Bingham, Fortuna, Commercial, Niagara, Dalton & Lark, and others.

The Jackling Enterprise

In a factual recording such as this chapter on Utah mining history there can be little reference to the many who contributed to the State's mining progress. However, so great was the influence of one man upon mining throughout the world, having as his inspiration the low grade disseminated copper ores of Bingham, that no reference to mining in the state or anywhere else could be made without a brief outline of his contribution, called "The greatest single example in world history of Man's collaboration with nature's resources, not necessarily for, but unquestionably to the benefit of human kind" on the copyrighted plate appearing on the front cover of Utah Centennial Issue of the Mining and Contracting Review for July 31, 1947

It matters not if it was the late Colonel E. A. Wall who "first noted copper discoloration on a hillside" at Bingham in 1887. Nor is it necessary to recount here the various versions from various sources of the controversy between factions as to who was the first to appreciate the possibilities of the low grade disseminated porphyry copper ore of the Bingham District. It is sufficient to say that Daniel C. Jackling, intrigued by the same ore while still busy with gold milling problems at Mercur (see section on Gold), was the engineer and metallurgist who wrote the report on the possibilities of the deposit that is to this day considered to be the most extraordinary example of vision and engineering imagination ("imagineering" as it is called today) in the annals of professional reporting; that it was Jackling who obtained initial financial backing (despite the jeers from many sources) to start the venture; that it was his perseverance that saved the venture on more than a few occasions; that it was his leadership and all-inclusive executive ability that made it possible to build up the staff who did the "impossible" and who subsequently carried the Utah Copper technique to all points of the globe; and, that the record of Colonel Wall's plans and his subsequent disposition of his holdings in the venture indicate that he had no conception of what the future held in store for the Utah Copper Mine. The reader is referred to the book, The Porphyry Coppers for an authentic story of the beginnings of Utah Copper and credit to those who stood by Jackling when most needed in the early days of the enterprise. For a slightly different version one may read T. A. Rickard's chapter on "The Great Salt Lake" in his A History o f American Mining.

The Jackling Enterprise started a new era in copper mining throughout the world because of its open cut mining and treating of very large tonnages of low grade ores, in which he was later assisted greatly by the application of the flotation process for concentrating low grade ore; in 1916.

It was in 1903 that the Utah Copper Company was formed and the first gravity concentrator was constructed at Copperton; 1905, when construction on the Magna Concentrator was started; 1906, when the Garfield Smelter was completed to smelt the concentrates from the Magna mill; and 1910, when the Utah Copper Company finally acquired the Arthur Mill in the consolidation with the Boston Consolidated Mining Company, a move initiated in 1908.

It is the consensus of all familiar with the Jackling Enterprise that the action of the Kennecott Copper Corporation, subsequent owners of the Utah Copper operations, in submerging the Jackling tradition and the same Utah Copper Company in recent years created deep indignation in the State of Utah and in the hearts of all who have remained loyal to Colonel Jackling. To most citizens of Utah it appears to have been an outstanding example of corporate stupidity.

Meantime, other contributors of copper in addition to the Utah Copper Company and the Utah Consolidated Mining Company, operator of the Highland Boy mine, entered the Bingham District. The present United States Smelting Refining and Mining Company, an outgrowth of the United States Mining Company which originated in 1899, is largely a consolidation of such mines as the Niagara, Fortuna, Commercial, Dalton & Lark, Yampa, Montana Bingham and other early copper producers. The Utah Apex Mining Company, formed in 1902 to operate the Utah Apex Mine (now merged with the Highland Boy properties and part of the National Tunnel & Mines Company, subsidiary of the International Smelting & Refining Company) was the winner in 1923 of the famous apex suit involving the ore bodies of the Utah Apex and Highland Boy mines. The latter mine became the property of the International Smelting & Refining Company as a result of the latter company having "backed" the loser financially. The original Ohio Copper Company (now the Ohio Copper Company of Utah) began its activities in 1912. The Bingham-New Haven Copper & Gold Mining Company, later the Utah Metal & Tunnel Company, now part of the National Tunnel & Mines Company, started mining in 1902. The Bingham Mines Company, since absorbed by the United States Smelting interests, began operating in 1908 and later acquired the Montana-Bingham properties.

Copper in The Tintic District

The mining of copper ore in appreciable amount in the Tintic District originated in 1869, mostly in the Copperopolis and Mammoth properties at Mammoth but the metal caused trouble due to the metallurgical difficulties of the times. Later contributors to the district's copper output were the Centennial-Eureka, Eagle & Blue Bell and Bullion-Beck, all since acquired at different times by the United States Smelting Refining & Mining Company; the Iron Blossom (now Tintic Standard); the Eureka Hill (now Chief Consolidated); the Grand Central (acquired later from Chief Consolidated by the American Smelting & Refining Company); and the Dragon mines.

In 1900 the value of annual copper output from the Tintic District reached a million dollars, in 1912 more than two million and again exceeded that value in 1917. In recent years the principal copper producers in the Tintic District have been the Chief Consolidated Mining Company; Tintic Standard Mining Company; Eureka Lilly Consolidated and Colorado Consolidated Mines (Tintic Standard).

From 1869 through 1947 the Tintic District has produced 121,321 tons of copper metal.

Copper in the Park City District

No record of copper recovery in the Park City District was made until 1898 when about 25 tons of metal was produced, most probably from the old Valeo and Odin mines. Most

copper in the district is derived from the lead-silver ore, and modern production comes as byproduct from operations of the Silver King Coalition Mines, the New Park Mining Company, and the Park Utah Consolidated Mines Company.

From 1898 through 1947 the Park City District has produced 34,813 tons of copper metal.

Copper in Beaver County

Shipments of very small tonnages from Beaver County started about the year 1872, but it was 1905 before the value of output reached $500,000. In 1906 and 1907 copper valued at more than a million dollars per year was produced, but it was the war year of 1917 before this record was again attained. Since then copper output from the district has been almost negligible. The principal copper producers of the district were the Cactus, Imperial, O. K., Old Harrington, Old Hickory, Montreal and Horn Silver mines.

Total value of copper produced in Beaver County probably has not exceeded $10,000,000 from 1860 to date.

Copper in the Ophir District

Copper recovery was first reported from the Ophir District in 1891, and the principal sources of the district's comparatively small output were the Hidden Treasure and Eureka-Ophir mines; some from the recent operation of the Hidden Treasure Mine by the United States Smelting Refining & Mining Company.

Copper in the Lucin District

Since World War I there has been no copper activity in the Lucin District. Small beginnings were made in 1868, and most of the district's less than 10,000 tons total output of copper metal was produced from 1905 to the end of World War I. The district is best remembered for the activities of the Salt Lake Copper Company, at the Copper Mountain Mine, which ended in financial failure during the 1890s.

Copper in Big and Little Cottonwood Districts

Although the Big and Little Cottonwood districts are reached through canyons miles apart, the United States Geological Survey records of their production have been combined.

The combined copper output of both districts probably totals about 10 tons of metal; no report of which appears in the records until 1902, although other mining started before 1870. At no time has the value of copper output per annum reached $500,000.

The principal copper producing mines in the Little Cottonwood or Alta District have been the Emma, Columbus-Rexall, Alta Cons, South Hecla, Michigan-Utah and Wasatch mines.

In Big Cottonwood most of the copper has come from the Cardiff, Big Cottonwood, Prince of Wales, and Woodlawn mines.

Silver

The all-time production of silver in Utah, second only to copper in value, through 1947 totaled 726,587,5 ounces, valued at $530,827,012. Almost all this silver has been the by-product of complex ores carrying the base metals.

The principal silver producing districts have been the Tintic, Park City, Bingham, Beaver County, Ophir, and Big and Little Cottonwood, American Fork, and Silver Reef.

It appears that the earliest discovery of silver in Utah was made in Alta in 1868, followed by the Park City and Silver Reef finds in 1869, and the American Fork and Mercur discoveries in 1870. Discovery in Tintic was made at about this time.

Silver in the Tintic District

From 1869 through 1947 silver has been the most important metal produced in the matter of value in the Tintic District. During that period the output of the district has totaled 260,773,729 ounces.

Early silver producing mines were the Scotia, Swansea, Showers, Crimson-Mammoth, Sunbeam, and Eagle. After 1900 important additional producers included the Eagle & Blue Bell, Centennial-Eureka, Victoria (all now owned by the United States Smelting Refining and Mining Company), the Grand Central (now owned by the American Smelting & Refining Company), the Chief Cons., Iron Blossom, Iron King (both now part of Tintic Standard). Later came the Tintic Standard and North Lily (International Smelting & Refining Company) mines.

Today consolidations in the Tintic District have reduced the number of major operating companies to the Chief Consolidated Mining Company, Tintic Standard Mining Company, the United States Smelting Refining and Mining Company, Mammoth Mining Company and the North Lily Mining Company.

Silver in the Park City District

As in the case of the Tintic District, silver has produced the highest financial reward in the Park City District. Most of the silver is a byproduct from the treatment of complex ores. Silver output since 1870 through 1947 totaled 235,979,700 ounces.

Although silver was discovered in 1869 by Rufus Walker, followed by others, it was not until 1872 that Rector Steen, John Kain, and one McDowell discovered rich silver ore sticking out of the ground. This was the origin of the fabulous Ontario Mine, sold in 1872 to Haggin, Tevis and George Hearst for $27,000. It was this property that contributed to the Hearst fortune, and up to the time of its acquisition in 1943, after long idleness, by the Park Utah Consolidated Mines Company, to be pumped out primarily for the recovery of the zinc ore left in the ground during the early days when zinc meant a smelter penalty, it had paid more than $14,000,000 in dividends.

Although the Buckeye, Flagstaff, McHenry, Pinon, and Walker & Webster properties shipped appreciable tonnages, it was the discovery of the Ontario mine that accelerated discovery and development in the Park City District. All this led to the later opening of the Daly, Anchor, Alliance, Mayflower, Daly-West, Silver King, Crescent, Daly-Judge, California-Comstock, Naildriver, American Flag, Kearns-Kieth, Keystone, and Quincy properties.

The Park Utah mine at Keetley was developed after World War I; the Park City Consolidated, Park City, in the late 1920s; and the New Park became a major shipper in 1939.

Consolidations have reduced the major producing companies in the Park City District to the Silver King Coalition Mines Company, Park Utah Consolidated Mines Company, New Park Mining Company, and Park City Consolidated Mining Company.

Silver in the Bingham District

From 1865 to 1873 silver contributed the most in dollar value from mining in the Bingham District, losing this leadership to lead for the first time in 1875. The recorded production of silver from the district from 1865 through 1947 totaled 147,952,523 ounces.

It would be difficult to exclude some silver from any of the Bingham properties enumerated under the section on copper. The United States Smelting Refining and Mining Company operations lead in silver output.

Silver in Beaver County

Discovered in 1875, the famous Horn Silver Mine was a heavy producer of silver for ten years; has produced intermittently ever since, bringing its total production of silver and lead to more than $20,000,000. The mine is credited with having paid almost $7,000,000 in dividends. Another important Beaver County silver producer was the Beaver Carbonate Mine, discovered in 1878. It produced silver-lead ore valued at about $ 1,000,000. The Cactus Mine produced somewhat less silver than the Beaver Carbonate.

The recorded total value of silver produced in Beaver County from 1860 through 1947 approximates $25,000,000.

Today the district is the scene of much investigation and exploration but of current minor importance from the standpoint of production.

Silver in the Ophir District

According to the Utah Mining Gazette of January 7, 1874, General Connor's soldiers discovered lead outcrops at Ophir in 1865; outcrops which led to the development of the Hidden Treasure silver-lead mine, but it appears that the Zella operations and the Pioneer Mill, built by the Walker Brothers in 1871 to treat Zella ore, were not only the first systematic efforts in the Ophir District but also the foundation of the Walker fortune. The Hidden Treasure Mine, now owned by the United States Smelting Refining and Mining Company, was the heaviest silver producer in the district, and although closed at this writing, was operated just a few years ago. Other early silver producers in the Ophir District were the Kearsage, Chicago, Mono, Douglas, Miner's Delight, and Wild Delirium mines. The Ophir Hill Consolidated and Lion Hill Consolidated mine operations, enterprises of the late Senator William A. Clark of Butte fame, became producers after the turn of the century, and continued so for about twenty-five years.

Today the Lion Hill region of the Ophir District is being explored by the International Smelting & Refining Company.

Unfortunately the early records of the United States Geological Survey did not separate the metal production of the Ophir District from that of the Stockton District. The latter district has been revived during the past twenty years through the reopening of the Honorine and Calumet mines by the Combined Metals Reduction Company.

The early day producers of silver in the Stockton District, where again General Connor's soldiers are credited with the discovery of silver in 1864, were the Silver King (credited with the first shipment of silver bearing lead ore from Utah), Galena King, the Honorine, the Quandary, and First National.

The value of silver produced in the Ophir and Stockton areas from 1870 through 1947 is estimated at $15,000,000.

Silver in Big and Little Cottonwood Districts

Silver was discovered in the Little Cottonwood District in 1864 but it was not until 1868 that the famous Emma Mine was located at Alta, followed by the Flagstaff in 1869. The next dozen years were the most productive and most wasteful in Alta's history. The mines contributing the most silver during this period were the Emma, Flagstaff, City Rocks, North Star, Nabob, Vallejo, and Toledo. Later producers have been the Columbus Cons, Michigan-Utah, South Hecla, Wasatch, Columbus-Rexall, Morning Star, and Sells mines.

For all-time production, the heaviest silver producers in Alta have been the Emma, Flagstaff, Vallejo, Columbus Cons, and South Hecla.

Alta has not been the scene of large production for more than twenty five years.

The silver producing mines of the early days in the Big Cottonwood District, established in 1870, were the Maxfield, Prince of Wales, Richmond, and Ophir. The Cardiff mine started its heaviest production in 1914, continuing dividend payments until slightly more than $ 1,000,000 had been paid by the end of 1927- Production since has been slight and the district is now the scene of search for new ore bodies.

The value of silver produced in the Big and Little Cottonwood districts from 1868 through 1947 is estimated at between $15,000,000 and $20,000,000.

Silver in the American Fork District

In the American Fork District, a comparatively short-lived district from the standpoint of major production of silver, the Miller Mine, discovered in 1870, was the big producer. Others have been the Wild Dutchman, Pittsburg, Bellerophon, Yankee, Silver Bell, Whirlwind, and Milkmaid. During the ten year period, 1870-1880, the American Fork District produced silver valued at $1,683,542. Since then annual production of silver has not been in excess of $100,000 (recorded) and except for exploration and minor leasing operations, the district is dormant.

Silver in the Silver Reef District

The Silver Reef District, where silver was discovered in 1869, was comparatively short-lived. During the period of recorded production of silver, 1875-1890, the value of silver output totaled $7,775,oo8. Since then it is doubtful if it has totaled $250,000, despite the attempt during the 19206 of the American Smelting & Refining Company to revive the Silver Reef.

The most productive mines in the Silver Reef District were the Barbee & Walker, Pinkham & Dodge, Leeds, Thompson & McNally, Gisborn, Silver Flat, Manhattan, Tecumseh, Buckeye, Last Chance, California, Toquerville, Dixie and Maggie Jane.

Lead

As in the case of copper, the Bingham District leads in Utah's all-time output of lead, with the Park City and Tintic districts, second and third, respectively, and Beaver County fourth. The Ophir and Stockton districts, and Big and Little Cottonwood districts rank next.

Bingham Lead Production

Lead produced in the Bingham District from the record, 1869 through 1947, has totaled 1,588,237 short tons of the metal. The first year in which the value of lead output from Bingham exceeded $1,000,000 was 1875, when it reached $1,566,000, although lead had been produced since 1865. Again in 1876 and 1877 the value exceeded $1,000,000, but did not reach that figure again until 1891, when the total value was $1,142,209, only to again drop off to a low of $94,000 in 1903. In 1906 the value was $1,104,259, and except for the year 1908, has steadily mounted ever since.

According to the United States Geological Survey, the all-time lead producing mines of Bingham, in the order of frequency of shipments have been the Bingham-New Haven, New England, Dalton & Lark, Utah Apex, Niagara, Highland Boy, Old Jordan, Queen, Montezuma, Utah Bingham, Story, Commercial, Butler-Liberal, Fortuna, Yosemite, Utah Copper, Phoenix, Massasott, Winnamuck, Last Chance, Central Standard, Yampa, Bingham Cons, Cluster, Mystic Shrine, United Bingham, Bingham Butte, Conglomerate, Ute Copper, Petro, Green Grove, Ivanhoe, Bingham Group, Boston Cons, Utah Metal, Bingham Mines, Montana-Bingham, Brooklyn, Julia Dean and St. James, most of which have been swallowed up by the consolidations outlined in the section of Copper in Bingham.

Later lead producers are the Butterfield operations of the Combined Metals Reduction Company, and the Bingham Prospect in the Lark area, now part of the United States Smelting Refining and Mining Company, producers of the major proportion of lead from the district.

Lead in the Park City District

Second to Bingham in all-time output of lead in Utah, the Park City District has produced 1,164,742 tons of the metal from 1870 through 1947. The history of lead in the district is similar to that of silver because of the close association of these metals in the ores in the area. The higher grade ores usually contain much lead and silver with some copper and gold. The lower grade ores were characterized by the presence of zinc, the early troublemaker because of separation troubles. The list of early shippers of lead and today's operating companies is similar to that for silver.

Lead in the Tintic District

History credits Steven Moore with having made the first lead discovery, the Sunbeam claim in the Tintic District in 1870. According to the records about $500,000 covered the value of lead shipped during the period 1870 to 1886. From then until 1894 the reported annual value of lead output did not reach $50,000. It reached a little more than a million dollars in 1898, and except for "off" years has continued to be of major importance to the district.

From 1870 through 1947 the Tintic District has produced 932,382 tons of lead metal.

For many years the mines that shipped lead most steadily were the Bullion Beck, Centennial-Eureka, and Eagle & Blue Bell, now all owned by the United States Smelting Refining & Mining Company; the Chief, Eureka Hill, Gemini, Plutus and Ridge & Valley, currently owned by the Chief Consolidated Mining Company; the Grand Central, lately owned by the American Smelting & Refining Company; the Colorado Cons, Iron Blossom, and Tintic Standard, of which the first two listed were acquired later by the Tintic Standard Mining Company; the Godiva; the Dragon, May Day, Mountain View, Uncle Sam and Yankee Cons, all of which later became the property of or under the control of the International Smelting & Refining Company; Sioux; Bowers, Gold Chain; South Swansea; Crown Point; and Lower Mammoth.

The North Lily mine, owned by International Smelting & Refining Company, which started shipping in the mid-1920s, is the only really major new lead producer in the district. Other new shippers of lead are the Eureka Lilly and Eureka Bullion mines.

Lead in Beaver County

According to Eissler's work on Metallurgy, a "Mormon" named Isaac Grundy melted lead ore for lead bullets in Beaver County in 1854, but it is said that the Rollins lead mine was the first mine in Utah.

The years 1882 and 1883, during which the value of lead mined reached $2,225,775 and $1,298,471, respectively, saw the largest output of the metal in the history of the district. This was during the prosperous years of the famous Horn Silver mine, discovered in 1875-but not so prosperous after 1885. As in the case of silver, the Beaver Carbonate mine, opened in 1878, contributed much lead. The O. K. mine also produced lead, mostly during the periods 1900-1901; and 1906-1907.

Other lead shipping mines have been the Moscow, Harrington, King David, Mammoth, Cedar Talisman, and Queen.

Estimated total value of lead from the district is $30,000,000.

In 1947 and since, the principal activity in Beaver County has been a revival of the Horn Silver properties and others controlled by the Tintic Lead Company.

Lead in the Ophir District

The history of lead, the most important metal produced in the Ophir district, has been much the same as for silver, insofar as the mines are concerned. The same is true for the Stockton District.

The value of lead output from the Ophir and Stockton districts is estimated at $25,000,000.

Lead in Big and Little Cottonwood

As in the Ophir District, lead and its importance in the Big and Little Cottonwood districts is similar to that in the matter of mines and production.

The combined value of lead shipped from the two districts approximates $25,000,000.

Gold

Utah has had some eleven or twelve gold producing mines from which the principal output was gold ore. In 1947 no such mine is in operation and the major gold producer is the Utah Copper open cut mine because of its large tonnage of copper ore containing small gold values. The annual output of gold from this mine alone exceeds former yearly production from the entire state.

Because of this by-product gold the Bingham District leads the state in total output, with the Tintic District next; the now no longer productive but very important Mercur District, third; and the Park City District fourth.

Gold first became important in Utah during the late 1890s, in the Tintic District, which, until Bingham forged ahead during World War I, was the leading gold producing district of the state. Prior to 1920 the maximum output of gold in Utah was reached in 1906 when its value reached $5,218,386. The Mercur, Tintic, and Bingham districts were the major contributors that year. With the decline in productiveness at Mercur there was a dropping off until the increasing tonnage at Bingham had its effect. In 1947 this district alone produced 381,000 ounces of gold valued at $13,335000. Utah's recorded output of gold through 1947 has been 10,996,076 ounces, valued at $287,187,525.

Gold in the Bingham District

Although gold was a major factor in the start of mining in Bingham, the value of annual production did not reach a million dollars until 1904. The early producers of gold from siliceous ores were the Old Jordan, Parnall, Niagara, Utah Apex, Buffalo, Gold and Silver and Utah Bingham.

Almost all the mines hereinbefore listed as copper, silver and lead producers have or continue to contribute gold to the district's total, because the crude ores of the district all contain gold, except possibly some crude lead-zinc ore.

The total recorded gold production of the Bingham District through 1947 has been 6,225,029 ounces.

Gold In the Tintic District

Although the Mammoth and Wyoming (Eagle & Blue Bell) mines produced gold in the early 1870s, gold did not take on major importance in the Tintic District until the late 1890s, reaching its first annual value of more than a million dollars in 1900 when production totaled 75,355 ounces valued at $1,557,726.

The principal gold producing mines in the Tintic District, with the Mammoth leading, have been the Mammoth, Grand Central, Eureka Hill, Eagle & Blue Bell, Chief, Centennial-Eureka, Victoria, Dragon, Iron Blossom, Swansea, South Swansea, and Iron King.

Most of the other mines hereinbefore listed under silver and lead producers have contributed gold because of its presence in the ores of the district.

Of later vintage in gold shipments are the Eureka-Standard (Tintic Standard controlled), now idle; the North Lily mine of the International Smelting & Refining Company; the Eureka Bullion and Eureka Lily mines.

The Tintic District's recorded gold production through 1947 has been 2,603,698 ounces.

Gold in the Mercur District

Although the Mercur District has not been an important contributor to Utah's gold production since 1917, despite continued courageous attempts to revive the district, its history, output, dividends and metallurgical significance are highly important.

Strangely enough the district got its name because in 1879 the first discovery was believed to be cinnabar by its locator, who called his claim by the name Mercur, the German for Mercury. Gold was discovered in 1883, following earlier flurries of silver mining which petered out to such an extent that the total recorded silver production of the camp probably has not amounted to more than $100,000.

The leading gold mines of the Mercur District were the Delamar, Geyser-Marion, Mercur, Sacramento, La Cigale, Overland, Sunshine, and Daisey. According to the United States Geological Survey, these mines produced ore valued at $19,034,984 during the period 1890-1917. And reported dividends from the Delamar, Mercur, Geyser-Marion, Sacramento, and Sunshine operations totaled $3,881,323.

After the original Mercur amalgamation mill, built in 1890, demonstrated the process to be non-applicable to Mercur gold ores, the cyanide process was introduced in 1892. Next the Marion mill, built originally to treat silver ore in 1872, which had been converted to an amalgamation plant in 1889, was changed to employ cyanide in 1893. In 1897 this mill was combined with the. Geyser mill of 1894 construction. The Sacramento cyanide mill was built in 1895.

Notable recovery failures were registered by the Sunshine (1895), Overland (1898), and the Daisy (1897) cyanide plants.

The Golden Gate mill was built during 1897 and 1898 under the direction of Daniel C. Jackling, who, as mentioned hereinbefore, already had his mind on the future Bingham enterprises. It was by far the largest cyanide plant in America, and, insofar as any records show, the first in the world to successfully roast ores of any kind prior to commercial scale leaching with cyanide solutions.

Inaugurated by Jackling in an experimental plant, many new departures were afterwards applied in the operation of the Golden Gate mill until the profitable ores of the Mercur and Golden Gate mines were exhausted.

The first of these departures was the calcining of semi-oxidized and somewhat carbonaceous clays so they could be leached by gravity percolation, but more outstandingly the roasting of so-called base ores containing considerable percentages of arsenic and other sulphides; as much as 1.5 percent or more of elemental arsenic combined in the form of mispickel, realgar and orpiment, the metallurgical conquering of which had baffled leading metallurgists of North America and Europe.

The Golden Gate mill was the first all-steel nonferrous metallurgical plant of any size in the world. It was the first nonferrous, or for that matter any kind of metallurgical plant, in the Western Hemisphere to use transmitted electrical energy as the source of motive power; the first in the world to use 3-phase induction motors throughout, the single-phase and multi-phase motor having been used only in small installations before.

Aside from the metallurgical innovations previously mentioned, zinc dust was first applied in the Golden Gate mill on a large scale for the precipitation of cyanide solutions; the first time that sodium cyanide, instead of potassium cyanide, was used as a solvent basis, the controlling reason for this being the larger percentage of the active radical cyanogen in sodium cyanide as compared to potassium cyanide, the first time anywhere, so far as is known, that a caustic soda or soda ash was used as a neutralizing agent to economize in alkaline cyanide consumption; the practice theretofore having been the use of calcium oxide (common lime); the first place in the world where zinc dust was being used, it was discovered that relatively pure lead-free zinc dust known as "blue powder" did not function effectively as a precipitant of gold from cyanide solutions, the remedy at Mercur having been the introduction of a small amount of lead to the zinc dust in suspension to form an electro couple. The almost infinitely small amount of lead was added in the form of lead acetate.

Probably the most outstanding of all the features incident to the building and operation of the Golden Gate Mill was the transmission of electric energy from the Provo River to Mercur, a distance of 43 miles, at 40,000 volts, transformed at one step at receiving station to a potential of 440 volts, the voltage for which all motors were designed.

Energy in commercial quantities had never before been transmitted as far; never before in commercial quantities transmitted at 40,000 volts or above, or for that matter, at more than a fraction of that potential; never before had a transformer been designed to convert 40,000 volts or other high potential alternating current to a safely usable potential of 440 volts or any other degree of high or low voltage transformation on a commercial scale. The generating and transmission system at Provo and thence to Mercur, and the use of power at the latter point, was the precursor of high potential electric transmission in the world in quantities of more than experimental volume of the nature of 10 or 20 horsepower.

This demonstration of long distance high potential transmission served as a basis for similar activities everywhere and led to an enlargement of the Telluride Power Company's system for the following fifteen years, when in 1912, the Telluride system was taken, in combination with others to form the basis for the present Utah Power & Electric Company, of which Jackling was the first president and operating administrator for a dozen years while the Utah Power & Light Company's Utah and Idaho power systems were being developed to practically their present status, including, during his tenure of office, the building, from Bear Lake to Salt Lake City, of the largest capacity and highest potential transmission system anywhere--in the form of parallel transmission lines 130 miles long, operating at 130,000 volts.

In a mining sense Mercur was the locale for another innovation in western mining. In 1897, the late Duncan MacVichie brought John McDonald and Joseph Hyland to Mercur to introduce the block caving system of mining, initiated on a small scale in the iron mines in Michigan, for the first time in the West, in the Delamar mine, along with Felix McDonald and Pat Hyland, all of them veteran iron miners. In 1903 it was John McDonald, at the direction of D. C. Jackling, who adapted for the first time, in Bingham, large-scale block caving, to low-grade copper porphyry ores, a system destined to be used in all the large copper porphyry properties where underground mining methods are employed wholly or in part, the forerunner of large scale block caving.

In 1900 the Mercur Gold Mining & Milling Company, of which the late John Dern was the first president, and the Delamar-Mercur Mines Company joined to form the Consolidated Mercur Mines Company. Since these operations ceased in 1913, Mercur has been almost dormant except for attempts to revive activity by the Snyder interests of Salt Lake City and one or two other organizations.

The total gold production of the Mercur District approximates $25,000,000.

Gold in the Park City District

In the Park City District gold has been less important, but output of the metal has increased because of a relatively new mine, the New Park. The first recorded shipments were made in 1881 when 34 ounces were shipped. In 1947 the district furnished 16,270 ounces to Utah's total.

Early shippers of siliceous gold ores were the Ontario, American Flag, and New York mines. Most the silver-lead mines of the district already listed produced some gold because of its presence in the complex ores of the area.

From 1881 through 1947 the Park City District has produced 578,076 ounces of gold.

Gold in the Gold Mountain District

The Gold Mountain District, long dormant, is worthy of mention because its gold (and silver) was derived exclusively from precious metal ores, unusual in Utah. From 1889 through 1947 the gold mined in the district, now inactive, totaled about $3,000,000 in value, mostly from the Annie Laurie and Sevier mines.

Gold in the Marysvale District

In the Marysvale area the Deertrail and Lucky Boy mines yielded gold from silver-lead-gold ores, starting in 1878. No authentic record exists on production prior to 1900, and since then not more than $800,000 in gold has been reported from the district, which is comparatively dormant.

Zinc

No official production of zinc in Utah was recorded until 1904, when that year's output was reported to have been valued at $16,979. As already stated, metallurgical problems involved in its separation from complex silver-lead-zinc ores and smelting penalties had caused zinc to be considered a detriment. Much of the early shipments were limited to lead-zinc sulphide ore which did not require concentration. The coming of the flotation process and differential flotation has made zinc important in Utah mining.

Although the Park City District was the early leader in Utah production of zinc, the Bingham District has long since gained first place, with Park City second and the Tintic District third. Beaver County, because of early production from the Horn Silver mine, stands fourth in all-time output. The Ophir and Stockton districts have contributed zinc, with the latter district's tonnage having increased during the past quarter century because of operations at Bauer.

From 1904 through 1947 Utah's recorded production of zinc has been 1,025,111 tons of metal valued at $147,823,341.

Zinc in the Bingham District

In 1909 the first reported zinc shipments from the Bingham District totaled 649,542 pounds, valued at $35,075. In 1916 the value of zinc shipped rose to $1,442,536. Early shipments of crude lead-zinc ore came principally from the old Winnamuck, Jumbo, Bingham, and Orleans mines. Zinc concentrates were derived from ores mined in the Commercial, Dalton & Lark, Old Jordan, Niagara, Utah Copper, New England, Utah Metal, Silver Shield, and Winnamuck mines.

Today the mines operated by the United States Smelting Refining & Mining Company (enumerated in previous sections) contribute the major part of Bingham's zinc output, with the Butterfield operations of the Combined Metals Reduction Company next since the suspension of operations by the National Tunnel & Mines Company.

From 1909 through 1947 the Bingham District has produced 498,197 tons of zinc metal.

Zinc in the Park City District

The records show the Park City District's first reported zinc shipments in 1905 to have produced 1,972,327 pounds, valued at $116,367, and it was 1910 before all-time value of recovered zinc reached a million dollars.

Early shipments of zinc concentrates came from ores mined in the Daly, Daly-Judge, and Daly West mines. The Daly-Judge and Daly West mines furnished most of the ores from which early lead-zinc concentrates were derived. Since the introduction of differential flotation the Silver King Coalition Mines, the Park Utah Consolidated properties and the New Park operations are the principal zinc producers. For a time, the Park City Consolidated Mine, currently scheduled for reopening, shipped zinc.

An interesting detail in 'connection with the comparatively recent acquisition of the Ontario Mine by the Park Utah Consolidated Mining Company, and its subsequent dewatering, was the avowed intent of first recovering the "worthless" zinc left in the ground by the pioneer miners.

From 1905 through 1947 the Park City District has produced 353,756 tons of zinc metal.

Zinc in the Tintic District

It was 1912 before any zinc production was recorded for the Tintic District-3,709,737 pounds, valued at $255,972. Early shippers of zinc ores were the Bullion Beck, Colorado, Chief, Empire, Gemini, Godiva, Iron Blossom, Lower Mammoth, May Day, Ridge & Valley, Sioux, and Uncle Sam mines. Lead-zinc ores have come principally from the Chief, May Day, North Lily, Ridge & Valley, Tintic Standard, and Uncle Sam mines. During World War II, below water level operations in the Chief mine contributed much needed zinc, and operations at still lower levels are now in progress.

From 1912 through 1947 the Tintic District produced 37,788 tons of zinc metal.

Zinc in Beaver County

Zinc shipments from the Beaver County District were first reported in 1904, with 332,924 pounds of metal valued at $16,979. Most of the crude zinc ore and zinc concentrates were shipped by the Horn Silver Mine, with some from the Cedar Talisman and Moscow mines. The zinc in lead-zinc ore has come from the Horn Silver Mine almost entirely.

The value of zinc produced in Beaver County from 1904 through 1947 is estimated at slightly more than $3,000,000.

Zinc in the Ophir District

Small shipments of zinc from the Ophir District were first reported in 1911. Early shippers were the Cliff and Hidden Treasure mines. Later the Lion Hill and Ophir Hill mines produced some zinc, and during World War II the United States Smelting Refining and Mining Company's operation of the Hidden Treasure mine added to the district's zinc output over the years.

In the adjoining Stockton District the Honorine and Calumet mines are the principal producers of zinc.

The estimated value of zinc produced from 1911 through 1947 in the Ophir and Stockton districts is approximately $3,000,000.

Iron Ore In Utah

According to history, the first iron works in Utah were established at Cedar City and Parawan in 1851. Today Cedar City is the nearest town to the principal sources of iron ore in the state. It was not until 1866 that W. L. Silver set up his iron works in Salt Lake City, the forerunner of Silver Brothers. The next important worker of iron to be established was the Provo Foundry & Machine Company, founded by Thomas F. Pierpont in 1895.

Statistics on iron ore produced in Utah before the year 1909 are vague and unreliable. In 1909 there were mined 34,634 gross tons, most of which was produced by the then Colorado Fuel & Iron Company. During the following year the tonnage almost doubled, with an output of 65,880 tons. In 1911 production dropped to 39,908 tons and it was 1916 before iron ore output exceeded that figure, with 45,514 tons. During the succeeding years annual output remained about the same until 1924, except for 1922 when the total fell to 15,000 tons.

With the blowing in of the Ironton pig iron plant of the Columbia Steel Company, the first iron blast furnace west of the Pueblo, Colorado, plant of the Colorado Fuel & Iron Company in 1924, production of iron ore from the Cedar City area of Iron County rose to 164,154 gross tons.

Much of the ore shipped to the Ironton blast furnace during its first few years of operation came from the properties of the Milner Corporation, although the Columbia Steel Company had its own mine, operated under the name of the Columbia Iron Mining Company.

During World War II, Utah's annual production of iron ore reached the one million ton mark, having totaled 922,959 gross tons in 1943; the next year 1,542,284; then 1,931,749 tons in 1945. Consumers of this tonnage were Geneva Mine, Carbon County, the Ironton plant of Columbia Steel Company, the wartime Geneva Steel plant, the Pueblo, Colorado plant of The Colorado Fuel & Iron Corporation, and the Fortuna, California steel plant of Kaiser, Incorporated.

When the United States Steel Corporation acquired the wartime Geneva plant, peacetime iron ore tonnage jumped to 2,741,000 gross tons in 1947, compared with 1,317,176 in 1946.

This in output has made Utah important in national production of iron ore. In 1947 Alabama, Georgia and Virginia produced 7,816,000 gross tons; New Jersey, New York and Pennsylvania, 3,872,000 tons; and the Lake Superior Region 76,699,000 tons.

There is much conflicting opinion concerning Utah's iron ore reserves. Estimates have risen from the 1920 figure of a scant 40 million tons to so high as the over-optimistic figures of 900 million tons, but conservative estimators now figure on Zoo million tons as a safe figure based upon drillings and exploration to date. Considering present plant capacity for pig iron production, the Ironton furnace of Columbia Steel (Geneva Steel Company) the Ironton furnace of Kaiser, Incorporated, and the Geneva Steel Company, known ore deposits appear sufficient to keep this capacity going for about 75 years.

The history of iron ore mining in Utah would not be complete without mention of L. F. Rains, the late Howard Gibbs, and Adrian C. Ellis. It was the determination of L. F. Rains that brought the first pig iron furnace to Utah.

Coal In Utah

As indicated in the opening paragraphs of this brief history of-mining in Utah, the ultimate value of the bituminous coal deposits may be said to rank next to that of the nonferrous metals-despite the prospect of atomic energy plants in the future. Short of presently inconceivable atomic energy plant simplification it appears that coal will continue always to be the source of direct heat and that Utah coal in particular is destined to yield great quantities of gas, oil and byproducts in the future.

Except for steadily diminishing reserves of high-grade coking coal, Utah's coal reserves are great. Great as these reserves may be, one must disregard some, optimistic estimates because of the distance back from the outcrop of some coal included. Some have set the figure at I 8o billion tons. The practical estimator figures the economically recoverable reserve at from 40 to 6o billion tons of bituminous coal. Be that as it may, if coal continues to be mined, when even 4o billion tons has been extracted the estimators will have passed into forgotten centuries.

The recorded production of coal since the beginning of its mining in Utah through 1947 has totaled 184,972,000 tons, with the 7,619,000 tons output in 1947, the highest in the history of the state. For the sake of conjecture, if one assumed an annual extraction of 10 million tons, the conservative estimate of 40 billions tons would last 4,000 years!

The discovery of Coal in Utah is credited to the 1776 expedition of Father Escalante and Father Dominguez into Utah, when they reached Spanish Fork Canyon and Utah Lake, but there is some disagreement as to where it was first mined and by whom. One version is that the followers of Brigham Young mined it first at Coalville in 1849, but that the first real mine was opened in 1854 at Wales, in Sanpete County. By 1865 early nonferrous smelting and some melting of iron ore created a demand for coal other than for domestic heating and lime burning. In 1875, following the discovery of coal in Huntington Canyon during the previous year, the Fairview Coal & Coke Company started operations at Connelsville. Four years later the Pleasant Valley mine was opened at Winter Quarters by the Pleasant Valley Coal Company. In 1882 the Mud Creek mine, not far from Winter Quarters started producing, and during the same year the Utah Fuel Company was formed to operate both the Pleasant Valley and Mud Creek properties. Next the Pleasant Valley Coal Company opened the Union Pacific mine at Scofield in 1884. Thus the three first "large scale" coal mines in the state were operated in or near the town of Scofield.

No new important mines were established until 1889 when the Clear Creek Mine, also not far from Scofield, was opened. This was followed in rapid succession by the opening of the Sunnyside No. 1 Mine east of Price by the Utah Fuel Company in 1890; the Castle Gate Mine by the Pleasant Valley Coal Company in 1890. Nine years later Sunnyside No. 2 was opened by the Utah Fuel Company.

Until 1906 the coal mines of Utah had all been controlled more or less by the railroads and when the Independent Coal & Coke Company started operations at Kenilworth, near Helper, the company was hailed by many as "a real independent, free of railroad domination."

In 1907 the first mines in the Hiawatha area, at West Hiawatha, was opened by the Consolidated Fuel Company; and the Mohrland Mine by the Castle Valley Coal Company. The next new opening was at Sego, east of Thompson, by the American Fuel Company in 1910. The year 1911 saw increased activity which included the start of the Panther Mine between Castle Gate and Helper by Frank Cameron; the Standard Mine in Spring Canyon by the Standard Coal Company; and the Black Hawk Mine near Hiawatha by the Black Hawk Coal Company. Following these new ventures came the Spring Canyon Mine of the Spring Canyon Coal Company in the, canyon of the same name west of Helper in 1912; the Cameron Mine of the Cameron Coal Company above Castle Gate in 1913; and the Carbon Mine of the Carbon Fuel Company at Rains in Spring Canyon in 1914.

The coming of World War I resulted in many new coal ventures and one important consolidation. Up until this time the Utah Fuel Company, having consolidated the Castle Gate, Clear Creek, Winter Quarters, Mud Creek and Sunnyside operations had been the only operator of several large mines. Formed in 1915, the United States Fuel Company (subsidiary of the United States Smelting Refining and Mining Company) acquired the West Hiawatha, Mohrland and Black Hawk mines in the Hiawatha area, and the Panther Mine near Castle Gate, all in 1916. It was in 1916 that the Rolapp interests acquired the Cameron Mine, now the Royal Mine.

Three new mines came into existence in 1917, the Latude of Liberty Fuel Company; the Peerless of Peerless Coal Company, both in Spring Canyon; and the Wattis Mine of the Wattis Coal Company, today the Lion Coal Corporation. In 1918 the Morton Mine of the Morton Coal Company was opened in Spring Canyon.

There were no more major ventures until 1924, when despite obvious over-capacity in the state because of the war expansion, the Consumers Mutual Mine (later called Blue Blaze) was opened by the Consumer Mutual Coal Company in an entirely new area, Gordon Creek Canyon, west of Helper. The following year the National Coal Company and the Sweet Coal Company followed suit in the same area.

At this time the first coal mining for the production of coke and byproducts in the making of pig iron in Utah was getting into its stride at Columbia, south of Sunnyside, where L. F. Rains and associates had opened the mine which became the property of the Columbia Steel Company, now part of the Geneva Steel Company, subsidiary of the United States Steel Corporation, war operator, and postwar purchaser of the wartime Geneva steel plant in Utah.

From 1925 to 1938 there were no major permanent additions to capacity in Utah. Although coal washing had been tried at Sego years before, it was not until 1938 that the first modern washer, sizer, and blender in the state was completed at Hiawatha by the United States Fuel Company. The Utah Fuel Company's plant designated to wash, size and blend coal from its various mines was completed at Castle Gate in 1939. Meanwhile the Independent Coal and Coke Company at Kenilworth had installed cleaning, sizing, and blending equipment but no washer.

In 1939, Terry McGowan, veteran coal mine superintendent, acquired the Blue Blaze mine in Gordon Creek Canyon and in 1941 the Hudson Coal Company took over the Sweet properties in the same field.

The Horse Canyon mine, south of Columbia, was opened in 1942 by the Columbia Steel Company for the Government to furnish coal for the by-product coke plant of the Geneva Steel Company. In 1946 the Geneva Steel Company acquired the Horse Canyon mine from the Government.

The establishment of the World War II steel plant at Fontana, California, created additional demand for coking coal and as a result the Utah Fuel Company entered into a working agreement with the Kaiser interests providing for coking coal from the Sunnyside properties of the Utah Fuel Company. This arrangement has continued with the peacetime operation of the Fontana plant. The Kaiser interests maintain their own staff of operating officials.

Industrial expansion in Utah and on the Pacific Coast serves to offset some of the pessimism concerning the coal industry's future in Utah. Currently foreign export keeps production at maximum capacity, with some fear expressed that more consolidations will be in order if the industry is to weather predicted lean years and ever increasing wage scales, material costs, suspensions due to strikes, and conversion of railroad power to oil burning locomotives and diesel locomotives, and municipal opposition to coal smoke.

Utah coal producers have never shown much interest in the processing of oil, gasoline, and by-products from coal, but if and when the day comes that the extraction of liquid fuel becomes economically necessary, the coals of Utah promise much because of their proved high content of oil, good quality by-product fuel and the ingredients for many chemical by-products. An approximate estimate of the value of coal produced in Utah from 1870 through 1947 is $452,000,000.

In 1899 the value almost reached the million dollar mark with $997,000; in 1900 the mark was passed with $i,448,000. Each year thereafter, with few' exceptions showed steady gains and in 1917, our first World War I year, the value jumped to $8,531,000; in 1918 to $13,937,000; dropped off to $12,800,000 in 1919, only to rise to $19,000,000 in 1920, from which figure there was a steady decline to 1934 when a new low of $4,700,000 was recorded. In 1935 the figures started to climb again and in 1943 made a new record of $19,800,000, followed by $22,000,000 the next year, $22,800,000 in 1945, in 1946 it was $21,750,000 and for 1947 the estimated value exceeded $30 million.

Other Minerals and Materials

Alunite In Utah

Alunite, the potassium aluminum sulphate of the Marysvale District was the source of potassium sulphate and was the subject of experiment for the extraction of alumina for aluminum during World War 1.

Again during World War II alunite was subjected to the Kalunite process at Salt Lake City and several shipments of the resultant alumina were made to a refinery at Tacoma, Washington, where it was refined to aluminum metal.

The Kalunite process for alumina and potassium sulphate as conducted in the Salt Lake plant was not economically successful for a number of reasons, principally the manner in which the plant was designed and "put together."

But the results proved that alumina can be extracted from alunite and that more favorable circumstances could quite possibly permit economical production, or if not, provide a source for emergency requirements of alumina and potassium sulphate.

Antimony In Utah

Since 1880 some antimony ores have been shipped from time to time from Garfield County and from the Brigham City area in Utah. In recent years there has been additional development in the Brigham City District.

Antimony occurs in Utah lead-silver-copper ores but statistics as to quantity and value are unobtainable.

Arsenic In Utah

Gold Hill in the Deep Creek District of western Utah has been the source of arsenic contained in arsenopyrite principally in wartime. During World War 11 the United States Smelting Refining and Mining Company produced arsenic at Gold Hill under a government contract. Some was produced during World War I.

Some arsenic is contained in Utah ores, notably in the Tintic District but it is of little importance to Utah mining.

Bismuth In Utah

According to A. Eilers in his notes on the occurrence of some rarer metals, each 100 tons of blister copper produced at the Garfield Smelter contains 6.1 pounds of bismuth.

The metal occurs in the ores of the Tintic and Clifton districts, and some shipments have been made from development work in the Alta District.

Up to the present time bismuth mining has been of little importance to Utah.

Elaterite--Ozokerite--Wurzilite in Utah

Unfortunately the records of production of Elaterite, Ozokerite, and Wurzilite, almost exclusively Utah products insofar as the United States, do not differentiate between the three minerals, all hydrocarbons.

Dana describes Elaterite as "an elastic bitumen, soft, elastic, sometimes like India-Rubber, occasionally hard and brittle." Ozokerite as "a mineral wax, essentially a paraffine." Wurzilite is mentioned by the U. S. Bureau of Mines in its reports on Gilsonite. The Soldier Summit area of Utah has been the principal source of Ozokerite.

No record of Ozokerite output can be found before 1920, and from then on the official records include Elaterite and Wurzilite in annual production figures of this rare insulating mineral.

As this is written, only one producer of Ozokerite in Utah, is listed, the Ozokerite Company, which produced 40 tons in 1947. The first record of production was in 1920, when 7,318 tons, valued at $110,400 was produced, the all-time record. Production has steadily declined except for occasional upsurges, as in 1924 with 628 tons and 1929 with 490 tons.

The combined tonnage and value of Ozokerite, Elaterite and Wurzilite from 1920 through 1947 is recorded as 11,400 tons, valued at $648,482, indicating an all-time average value of about $57 per ton. During the past few years the value per ton has been between $75 and $8o.

Gilsonite In Utah

Gilsonite, used in the arts, special varnishes and paints, in special insulators, and paving mixtures is almost exclusively a Utah product. It is hydrocarbon.

The first recorded mention of Gilsonite, called Uintahite and Uintaite by William P. Blake until S. H. Gilson of Salt Lake City first brought it to commercial notice and created from it certain insulating compounds, was in 1885. The first producing company was formed in Salt Lake City, and in 1889 its total shipments came to 1,500 tons. During the same year a new St. Louis company shipped 6oo tons. This company at one time registered the name "Gilsonite" as a trade mark.

The first Utah Gilsonite vein mined was called the "Carbon Lode." It was located near Fort Duchesne, about 100 miles northeast of the town of Price. During the I89os the twenty-mile Dragon-Rainbow-Pride of the West lode was developed by the Gilson Asphaltum Company, a subsidiary of the Barber Asphalt Company. At first the sacked ore (still sacked) was hauled by six and eight horse and mule teams over the 8,500 foot Baxter Pass to Fruita, Colorado, for shipment to St. Louis over the Denver & Rio Grande Railroad and its connections.

Increased demand for Gilsonite was responsible for the next step in providing for its transportation to market-the building of the Uintah Railway, a daring engineering feat. The narrow-gauge railroad, ultimately 65 miles long, was built from Mack, Colorado, to Atchee, Colorado, thence up and along the edges of sheer precipices on a 7 percent grade to the summit of Baxter Pass, down the other side into Uinta Basin to Dragon, its first Basin terminus. Shay engines were used from Atchee to the summit. The railroad started operating in 1905 and later was extended to Watson, and a branch to Rainbow, the center of operations of the Gilson Asphaltum Company. The railroad, now torn up, has given way to truck transportation to Craig and Fruita, Colorado, and Price, Provo and Salt Lake City, Utah.

Until comparatively recent years the mining of the vertical veins of Gilsonite was done on what some thought to be a "primitive" basis. There were reasons for this. From 1885 until a comparatively few years ago, absence of permissible explosives, small annual demand, and many miners skilled in the non-blasting system of mining, governed the methods employed in extracting the highly inflammable and explosive material. The extraction methods have been simple. Shallow shafts were sunk in the vertical or nearly vertical Gilsonite veins, bounded by extremely hard wall rock, and mining was (and still is in some properties) done by a sort of top-slice, underhand stope method, for 75 to 100 feet on either side of open shafts, without shaft pillars. As mining advanced the walls of the vein were (are) supported by horizontal stulls, the mined-out portion of the vein open clear to surface. In most cases access to the stopes was entirely by ladder; shafts were close together and access between them was possible only by climbing up and down the ramps created by mining operations. As depth was gained the miners had much open "country" above their heads.

Much of the ore was dug out with hand picks, later with pneumatic picks until permissible explosives were introduced. In general, relieving the pressure on the vein by channeling along the side walls facilitated extraction, in many cases caused the material to "bounce" out from the vein in chunks. All the ore was shoveled into sacks and hoisted for shipment in a chain sling.

In 1938 the Barber Asphalt Company moved its mining to the Bonanza Lode, with the center of operations to Bonanza, some 48 miles southwest of Vernal, Utah. Early in 1943 increased demand for Gilsonite and a shortage of skilled Gilsonite miners, who had entered the armed forces or war industries, caused the Barber Asphalt Company to discard the traditional method of mining and to introduce revolutionary methods in mining Gilsonite. The company adopted shrinkage stoping methods, compressed air drills and permissible explosives in breaking the ore. Shafts were equipped with steel headframes and steel shaft bins, electric hoists and air compressors. A cleaning and sizing plant was built and all product sacked after cleaning and sizing.

Tonnage per man was increased several times. Unfortunately, a serious fire and explosion (anticipated by many veteran Gilsonite producers) occurred on October 9, 1945, and since then the use of explosives has been prohibited, although carefully applied mechanization continues. Today the American Gilsonite Company, jointly owned by the Barber Asphalt and Standard Oil of California companies, operates the Bonanza properties, and is the biggest producer of Gilsonite in the United States.

Few, even citizens of Utah, realize the dollar magnitude of the work carried on for years in the "primitive" Gilsonite mines of Utah. The first reliable record of tonnage and its value is dated 1903. In that year a little more than 5,000 tons, valued at $188,357 were shipped. In 1910 output rose to 35,697 tons, worth $569,333, for the first annual value of more than half a million dollars. This was exceeded in 1912 with a total value of $633,069. The World War I years, 1916, 1917, and 1918 produced annual values of $633,440, $569,325, and $633,258 respectively. Annual value of production first reached the three-quarter million dollar mark in 1925, with $767,900. The million dollar range came in 1928 with $1,037,679, followed by $1,235920 in 1929. The value of Gilsonite mined in 1939 was $1,053,192, and in 1943 it reached $1,188,485. The figures for 1945, 1946, and 1947 were $1,250,546; $1,572,900; and $1,485,300, respectively, with the last two figures estimated. Tonnages in 1946 and 1947 rose to 74,900 and 70,735 tons, respectively, creating all-time records.

From 1903 through 1947 approximately 1,535,000 tons of Gilsonite valued at $30,000,000 was produced in Utah.

Demand for Gilsonite increases in time of war and during prosperous years.

Today the principal Utah producers of Gilsonite are, the American Gilsonite Company, American Asphalt Association, the Black Virgin Mine, the Ray Davis operations, Dragon Gilsonite Company, Raven Gilsonite Company, and the Utah Gilsonite Company.

Any review of the Gilsonite industry of Utah must mention Homer D. Ford, for many years in charge for the Barber Asphalt Company (Gilson Asphaltum Company), and Charles J. Neal of Vernal; and, of course, S. H. Gilson, already mentioned, after whom the unique mineral is named.

Mercury In Utah

Although the old gold camp of Mercur was so named because one prospector took it to be a quicksilver camp, Utah has produced very little quicksilver.

The main source of the metal was the gold mining operation of the Sacramento Gold Mining Company at Mercur, credited with a byproduct production of 3,338 flasks. The only other source of record was the Lucky Boy Mine in the Marysvale District, where 213 flasks were produced, during the period 1891-1897

No mercury output appears to have been recorded since 1907 except a few flasks. From 1881 through 1907 total production was 3,551 flasks valued at $139,000.

Molybdenum In Utah

Recognized for many years in small amounts in the Bingham, Beaver Lake, Clifton, Lucin and other districts of Utah, and despite small shipments from the Little Cottonwood District, molybdenum did not become important to the state until demand for its use in the steel industry, mostly as a substitute for Tungsten, and metallurgical advance made its recovery from the copper ore of the Utah Copper pit in Bingham economically advisable.

As a matter of fact, D. C. Jackling and his staff recognized the potential importance of molybdenum at the outset of Utah Copper operations but at the time the metal did not have commercial attractiveness.

Total domestic production of molybdenum in 1914 was only 1,297 pounds. During World War I total U. S. output reached 861,637 pounds, but in 1921 and 1922 no production was reported. In 1925 our national production passed the million pound figure for the first time, and has increased steadily since, particularly in time of war.

The product derived from the copper ore of the Utah Copper pit is Molybdenite, the disulfide of Molybdenum, which is 59.95 percent Molybdenum. In 1936 the Utah Copper Company began separating and recovering Molybdenite on a small commercial scale from copper concentrates before their shipment to the smelter. In 1937 the Utah Copper Company produced 8,187,615 pounds of Molybdenite concentrates containing 4,912,569 pounds of Molybdenum metal.

During the period 1937 through 1946--nine years--the copper ores of the Bingham pit yielded 139,000,000 pounds of Molybdenite, roughly 41,700 short tons of the metal!

Thus the Utah Copper Mine has placed Utah second only to Colorado in output of Molybdenum in the world during periods of high demand for the metal, and first in world production during periods when slack demand does not warrant capacity operation of the famous Climax Mine in Colorado. Since the Kennecott Copper Corporation no longer segregates production of metals from its various properties, current annual production of Molybdenite in Utah requires calculations involving past output in relation to total copper ore tonnages-all of which is left to the reader.

The U. S. Bureau of Mines does not segregate the value of Molybdenum metal produced by states or producers with the exception of the Climax Molybdenum Company in Colorado.

During the peak war year of 1942, U.S. sources produced 33,218 tons of Molybdenum metal, valued at $47,275,000.

The Utah Copper Mine is the only producer of Molybdenite of record in Utah.

Tungsten In Utah

Tungsten, so far of little economic importance to Utah, had long been known to occur in Box Elder County, in Tooele County and in the Little Cottonwood District when the advent of World War II and the fluorescent light for prospecting gave the metal momentary minor commercial value.

Scheelite, calcium tungstate, low in grade, was found in the Milford District of Beaver County, principally in the Old Hickory Mine, associated with copper, and small amounts of silver and gold.

During peak World War II demand for Tungsten, a 125-ton mill at Milford produced a total of not more than 2o tons of high grade concentrate varying between 55 and 65 percent tungstic oxide, but the postwar slump in price and demand stopped any production of record in the state.

It appears doubtful that Utah will ever be the source of much Tungsten, the bulk of which is produced domestically in California, Idaho, Nevada, and Colorado.

Vanadium--Uranium--Radium In Utah

In Utah Vanadium and Uranium (and Radium) are so closely associated because of the ores in which they occur.. In general, Utah Vanadium operations involve the recovery of Uranium and Radium in addition to the Vanadium, from the principal ore, Carnotite, potassium uranyl vanadinite.

According to the records Carnotite was first found in Utah in 19o9, with promising developments on the Kent and Bonanza claims located about ten miles from Bedrock in San Juan County. In 1911 the Utah Rare Metals Company shipped about 30 tons of Carnotite to domestic consumers and some concentrate to England; and in 1912, V. C. Ward shipped a carload of Carnotite from deposits near Richardson.

During World War I the demand for Vanadium increased, as did small shipments from scattered workings in southeastern Utah, and in 1920 the output of Carnotite in the Moab area increased. By 1924, the old United States Vanadium Company (no connection with the present company) acquired Carnotite claims on Dry Wash, 22 miles northeast of Monticello, from the Radium Corporation of Colorado, where a 20-ton mill was built. During the same year ore from Yellow Cat Wash, twenty miles east of Thompson, was shipped to Denver. During 1925 the Utah Vanadium Company, in conjunction with the Pittsburgh Radium Company, mined Carnotite on the Yellow Cat Wash, and in 1926, H. W. Balsley mined Carnotite in the Moab area for its Radium content. In addition the Utah Vanadium Corporation continued its Yellow Cat Wash operations, with the United States Company producing ore form Dry Valley, forty miles southeast of Moab.

Two new producers entered the Utah field in 1929. The Vanadium Alloys Corporation acquired the San Juan County holdings of the old United States Company, and the International Vanadium Mining Corporation (formed by B. F. and B. A. Pitts) purchased the Dry Valley properties of the Vanadium Alloys Corporation and built a mill. In 1930 the Dry Valley properties were operated by W. C. Patterson. During the following year H. W. Balsley of Moab sold some rich Carnotite ore for therapeutic and fertilizer purposes.

Following two years of depressed market for Carnotite ores, in 1933 activity increased in Utah, where Frank Silvery mined ore near Summit Point; Shumway Brothers near Blanding; and the Molybdenum Corporation of America acquired the Dry Valley claims from the International Vanadium Corporation. In 1934, H. W. Balsley's Yellow Circle Mining Company mined in the LaSal Mountains, sixteen miles from Moab, and J. W. Lewis made shipments from Grand County. In 1935, H. W. Balsley of Moab was Utah's principal shipper of Carnotite from 30 small mines under contract, and J. W. Lewis made shipments.

The year 1936 saw a quickening of activities in Utah, helped by the building of a new treatment plant of Uraven, Colorado, to which the small Utah producers could ship-built by the United States Vanadium Corporation, and Rare Metals, Incorporated, which entered the Moab area. During 1938 Utah production of Carnotite was scattered, with Harbo Mines at Cisco, contributing most of the state's total. In 1939 and 1940 the approach of war increased Utah output, with shippers sending their ore to the U. S. Vanadium Corporation's plant in Colorado, built in 1939. During this year the Shumway properties near Blanding led in production from Utah, with Harbo Mines, Cisco, contributing concentrates. Total Utah output of Vanadium metal was 113,813 pounds.

Censorship during World War II prevented any realistic estimate of how much Uranium, Vanadium or Radium was recovered from the mining of Carnotite or other ores in Utah, but records of previous years give clues if applied to total tonnages mined. Figures on U. S. production for one year reveal that 1439 tons of Carnotite yielded 52,695 pounds of Vanadium, 17,951 pounds of Uranium, and 2,716 milligrams of Radium. During the same year 74,229 tons of other ores yielded 86,817 pounds of Vanadium. The record for another year shows that from 1,708 tons of Carnotite there were produced 73,788 pounds of Vanadium, 20,764 pounds of Uranium and 3,141 milligrams of Radium, while 129,272 tons of other Vanadium ores yielded 1,012,337 pounds of Vanadium.

At this writing it would appear that the Radium content of Utah Carnotite has became less important than in the past because of the radioactive properties ;of the byproducts of Uranium and Thorium and the ability to activate other metals such as Cobalt for radioactive medical purposes.

During World War II the Monticello and Blanding plants were built by the government and operated amid much "hush" as to the type and quality of recovery made. As this is written the Atomic Energy Commission is preparing the Monticello plant for reopening in connection with the Commission's program instituted to encourage the finding of new Uranium bearing deposits and increasing production under guaranteed prices and operation of the plant, as well as a bonus for the discovery of entirely new rich non-Carnotite Uranium deposits. Whether the prices offered (already declared inadequate by the independent producers of Utah and Colorado) will stimulate production or aid the small miner remains to be seen.

Objective consideration of the Utah Vanadium-Uranium industry cannot lead to any conclusion but that the small independent miners have been pretty much at the mercy of the interests who have controlled private production in the United States. That this continued up until the time of the Atomic Energy Commission's 1948 incentive plan is evidenced by the following official U. S. Bureau of Mines Statement:

In 1939 the two outstanding producing companies (U. S.) agreed to share the world market because their output potential exceeded absorption capacity, without a major price reduction.

The relative domestic importance of Utah's Vanadium may be gauged by figures for 1943, during which the percentages of total U. S. production were: Utah 15; Colorado 74; Arizona 4; with the remaining 7 percent contributed by Idaho, Nevada and New Mexico. As for Uranium, the reader is left to his own calculations.

Miscellaneous Minerals In Utah

Rock Asphalt

In 1891 California rock asphalt was used to pave one short street in Salt Lake City. This stimulated interest in and the development of Utah deposits in 1892.

Early Utah production came from a limestone rock asphalt deposit in Spanish Fork Canyon. Despite high costs and inadequate equipment, some twenty Salt Lake City streets were paved with this material during the period 1892-igi6, after which the industry was dormant until the 1920s when it was revived at Sunnyside, but the promoters' progress in obtaining use of rock asphalt in many states was nullified by financial difficulties and extravagant management.

Again in 1930 the sandstone base rock asphalt operations at Sunnyside were revived by Howard C. Means, former Chief Engineer of the Utah State Road Commission. Costs were reduced and output increased during the years 1930 to 1947. In 1947, the operation's biggest years, circumstances beyond the control of Howard Means, involving shipment of inferior grade material, brought about a financial crisis, currently being adjusted through reorganization. .

There are numerous deposits of rock asphalt in Utah but the Sunnyside bed appears to be one of the thickest in the United States. The bed is known to be 1,285 feet thick, and contains about 10 percent asphaltum and go percent pure silica sand, giving a product having every void permeated with asphaltum. It is this characteristic that has made possible pavement and surfacing of uniform texture, without spots or irregularities-when properly prepared and of correct grade.

Statistics on output of rock asphalt are scanty. In 1946 and 1947 the Rock Asphalt Company of Utah produced 26,000 and 30,892 tons respectively.

Salt

Contrary to common impression, the production of salt from the brine, residual salt of Great Salt Lake, and rock in Utah has never reached the magnitude one would expect.

Production records from 1880 are available. Starting with 12,000 tons, valued at $60,000 in 1880, the annual value of production reached $102,000 in 1887; then dropping off until 1890 when it came to $126,000, followed by $265,000 in 1891, to $340,000 in 1892. It was 1901 before the value again passed the $300,000 figure, with at total of $326,016. In 1918 a new record of $580,375 was attained, a record not since equaled. Production in 1945 and 1946 was valued at $363,997 and $339,505, respectively, for 122,997 and 121,669 tons.

From 1880 through 1947, Utah produced 4,182,000 tons of salt valued at $15,650,000.

Gypsum

Although the records show that the commercial production began in 1890, these records are incomplete and misleading in that they do not give any clue to the current importance of the industry. This importance is indicated by the construction of two new plants in the Sigurd area and greatly increased annual output.

Although the Nephi Plaster Company was established in 1889 no gypsum records are listed until 1910, showing 46,2 79 tons valued at $149,089. Earlier records show only annual records of value running from $5,000 to $16,000 per year. Statistics show that the highest recorded value was reached in 1924 when 90,221 tons valued at $335,588 were produced. After this record was attained production declined steadily, and from 1931 to 1947 the annual of output remained below $100,000 per year.

It is impossible to present all-time tonnage and value figures on gypsum produced in Utah.

Sand and Gravel

Nature, with its sand and gravel debris, has made the sand and gravel industry important in Utah.

Starting with a modest 20,990 tons valued at $3,599, recorded for the year 1908, sand and gravel output more than doubled during the following year; first reached a value of more than $100,000 in 1918; has increased rapidly since; in 1920 some 519,000 tons worth $252,500; in 1923 the million ton mark was passed with 1,212,077 tons valued at $290,807; the two million ton figure was exceeded in 1936 when 2,267,808 tons worth $1,352,296 went to market; only once has the three million ton mark been exceeded, in 1942 with 3,874,841 tons valued at $2,079,302 established an all-time record. Recorded Utah output of sand and gravel from 1908 through 1946 has been 43,500,000 tons valued at $20,300,000

Stone

The first recorded statistics on the production of stone, building and crushed, for Utah are given for the year 1907. Obviously the available totals are entirely misleading because, if for no other reason, they omit the value of the granite and sandstone used in the building of many "Mormon" temples--particularly the temple in Salt Lake City--many buildings, and other structures.

The value of $338,328 is given for stone produced but there is no tonnage figure for 1907. The earliest production figure appearing is for 1919, which of course omits the volume of Little Cottonwood granite used in building the beautiful State Capitol in Salt Lake City. In 1920 the value of stone produced reached $509,740 for 304,290 tons. The record for value is shown to have been 1928 when 881,730 tons valued at $818,415 were produced, but the all-time tonnage record was made in 1940 with 1,024,660 tons worth $693,127.

The stone business has declined since 1940. The latest available figures are for 1945 when 515,400 tons valued at $318,254 were produced.

Any attempt at total stone output in Utah or its value would be highly inaccurate and misleading.

Lime

The records of the lime industry in Utah are also inadequate, starting so late as in 1904, with 19,000 tons but no value. In 1905 some 12,765 tons valued at $60,089 were produced. Not until 1920 did the value of lime produced exceed $100,000 with 9,797 tons worth $151,700. The half million dollar mark was passed in 1925 when 22,062 tons sold for $265,060. Only once has the annual value reached more than four hundred thousand dollars-in 1943 with 59,811 tons worth $407,753. There has since been a decline. In 1945 some 47,500 tons valued at $353,671 were produced. From 1905 through 1945 the total recorded output of lime in Utah was 985,000 tons valued at $7,750,000.

Fire Clay

Despite the importance of refractories and brick in Utah, the production records are so inadequate as to render comprehensive figures impossible.

The brick industry was started in 1895 by the Salt Lake Pressed Brick Company, now the Interstate Brick Company. The Utah Fire Clay Company established its clay products plant in Salt Lake City in 1904.

The meager statistics available for fire clay output extend only back to 1906. Output has varied from a minimum of 1,740 tons valued at $12,700 to a maximum of 20,441 tons worth $40,107 in 1940. Since then production has declined.

Manganese

Despite shipments of manganese ore from the La Sal District by the then Colorado Fuel & Iron Company as early as in 1901, the manganese industry in Utah has been largely one of wartime activity. Low grade ore from Juab County has been the principal product. During World War II production totaled about 300 tons in 1939; almost 6oo tons in 1941; a few tons in 1943; approximately 5,000 tons in 1945; and 8,000 tons in postwar 1946-all manganiferous type ore.

Fluorspar

Except for the World War II born effort of the Tintic Standard Mining Company to mine and treat fluorspar from the Cougar deposit reached by way of Milford, there has been no fluorspar industry worthy of note in Utah.

About 6,500 tons of medium grade fluorspar concentrate were produced from 1939 through 1946.

Phosphate

Potentially of possible importance to Utah, phosphate rock occurs in many localities in the state but to date has not been of commercial importance.

Sulphur

The surface deposit of sulphur at Sulphurdale was first exploited on a fair scale a quarter of a century ago but during the past few years has not warranted the publication of official statistics.

Petroleum

In reality there is no history of commercial production of petroleum in Utah. But the search for petroleum has brought about the discovery of helium, natural gas, and carbon dioxide gas on an important scale. Magnesium and potassium beds have been disclosed in the search.

Except for one or two temporary spurts of petroleum from drilling operations and in numerous reports of "showings" as this is written, commercial production of petroleum in Utah has yet to come.

In the past thirty to forty years there have been drilled about 130 dry wells; 40 with "oil showings;" 25 with "gas show;" a dozen gas producers; one helium well; five carbon dioxide gas producers; and six asphalt wells.

The largest "momentary" delivery of oil appears to have come from a well on the Cane Creek Anticline in Grand County. Also the Utah Southern well on the Salt Valley Anticline in Grand County. There have been spectacular but short-lived "blow-outs" of oil from wells in the Crescent Thompsons area of north Grand County.

During 1947 there was a great renewal of interest in the possibility of petroleum in commercial quantity in Utah. Responsible were the immense output from the Rangely, Colorado, pool, close by; the "scarcity of oil;" new concepts of Utah geology; improved geophysical methods; and drilling equipment capable of going far deeper than the depths accompanying previous "failures." Most of the drillings had reached a maximum Of 3,000 to 5,000 feet, with a few exceptions.

The consensus of oil geologists is that past failures have been due to drilling "off-formation." Many agree with Glen M. Ruby, prominent oil geologist in his warning (against "the mistakes of the past") that the apparent simple geologic structure of Southeastern Utah has led to mistakes in both structure and stratiography, and that Utah has a geology all its own-that it is a mistake to assume that its geology is like that of Wyoming, Oklahoma, or California.

One must await results from the drillings of many of the responsible producers and exploration companies currently obligated to drill in favorable locations with modern equipment capable of going to great depths.

Natural Gas

There are two natural gas producing areas in Utah. One is the Ashley Valley field near Vernal, where several shallow gas wells have produced for a good many years. The other is the Clay Basin field where a half dozen wells are producing, with more being drilled by a Utah utility company. This company continues the search for commercial natural gas in other parts of the state.

Helium

Drilled in the search for petroleum, the helium gas well near Woodside in Emery County is reserved for the United States. Government. Minute percentages of helium gas have been reported in other drillings but the absence of government reserve indicates insufficient. content for. the wells to be called helium producers.

Carbon Dioxide Gas

Oil exploration has resulted in the creation of one going area producing carbon dioxide gas for conversion to dry ice; the demand for which is expanding rapidly-and a new potential field: '

The first is on the Farnham Dome, near Wellington-in Carbon County, where since the late 1920s one concern . has been increasing -its capacity steadily. Statistics on carbon dioxide are obscure. However, some idea of Utah's dry ice output compared with -total i7: S. production (from gas and synthetic gas) may be gleaned from the fact that in 1946 the Utah producer manufactured 6,886 tons compared with the U. S. total of 644,592 tons. The 1946 Utah capacity of 25 tons per day was increased to 4o tons.

The other field, yielding heavy carbon dioxide gas, was drilled into west of Price in Carbon County. during 1947 and a large capacity plant is under consideration.

There have been showings of CO2 gas in numerous other drilling operations.

Magnesium

The United States Bureau of Mines and the then Secretary of the Interior, Harold L. Ickes, are the authority for the statement that the "Defense Well" of World War II, drilled in the Crescent area in , Grand County, revealed fabulously thick beds of magnesium and potassium salts. In addition, the well forced appreciable quantities of petroleum out through its casing, and although capped, oil continues to ooze through the capping. Other wells in the vicinity have shown definite magnesium and potassium content in standing liquid for many years. These wells have also spurted oil over the landscape from time to time.

The "Defense Well" was mysteriously capped and abandoned by the Government soon after the announcement proclaiming its fabulous discoveries.

Rarer Metals In Utah

Because the smelters render no reports on the rarer metals in the ores of Utah, it is not possible to give definite figures on production. However, A. Eilers, in his notes on rarer metals contained in the blister copper produced at the Garfield Smelter says that each 100 pounds of this blister copper contains 56 pounds of selenium; 5.4 pounds of tellurium; 0.342 ounce of platinum; and I.I83 milligrams of palladium.

###