Marine Biological Laboratory Received 66756 Accession No A^« 1952 Given By_ Place, Universitv of Wisconsin Press Maaison, Wisconsin " MINERAL NUTRITION OF PLANTS This Volume is Published in Celebration of the HUNDREDTH ANNIVERSARY of the Founding of the University of Wisconsin Contributors DANIEL I. ARNON O. BIDDULPH G. B. BODMAN DAMON BOYNTON T. C. BROYER HANS BURSTROM HARRY F. CLEMENTS L. A. DEAN JACKSON B. HESTER HANS JENNY CHARLES E. KELLOGG W. F. LOEHWING C. EDMUND MARSHALL A. G. NORMAN A. C. ORVEDAL ROY OVERSTREET J. B. PAGE L. A. RICHARDS ROBERT A. STEINBERG EMIL TRUOG C. H. WADLEIGH ROBERT B. WITHROW MINERAL NUTRITION OF PLANTS 71 Edited by EMIL TRUOG 1951 THE UNIVERSITY OF WISCONSIN PRESS Copyright ig^i by The Regents of the University of Wisconsin Copyright, Canada, 795/ Distributed in Canada by Burns & MacEachern, Toronto PRINTED IN THE UNITED STATES OF AMERICA BY THE WILLIAM BYRD PRESS, INCORPORATED, RICHMOND, VIRGINIA W.A.S. PUBLISHER'S NOTE For bibliographical reasons, Professor Emil Truog, chairman of the committee responsible for this book, has been designated editor. Listing volumes of essays written under separate author- ship presents problems to bibliographers which do not easily lend themselves to practical solution. The Press, therefore, feels that scholars will be grateful for a simple entry under which this book may appear in files, catalogues, and bibliographies. Program Committee DAMON BOYNTON American Society of Horticultural Science JACKSON B. HESTER Fertilizer Section, American Chemical Society W. F. LOEHWING American Society of Plant Physiologists D. D. LONG Plant Food Research Committee, National Fertilizer Association M. H. McVICKAR National Fertilizer Association A. G. NORMAN Chemical Corps, War Department ROBERT A. STEINBERG Botanical Society of America EMILTRUOG, Chairman Soil Science Society of America G. W. VOLK American Society of Agronomy C. H. WADLEIGH American Society of Plant Psysiologists ROBERT B. WITHROW Smithsonian Institution Preface f V Iineral nutrition of plants is a subject of tremendous interest and importance to many people. The plant physiologist must, of course, devote much of his attention to this field; the agronomist, horticulturist, and forester meet problems almost daily which require for their solution specific knowledge dealing with the mineral nutri- tion of many kinds of plants; and lastly, the fertilizer manufacturer, who is called upon to supply the needed mineral nutrients when they are lacking in soils, must, through his technical expert, keep informed of the latest findings in this field if his enterprise is to attain and maintain a forefront position. Because these people representing several different but related lines of activity seldom have a suitable opportunity as a group to discuss those phases of the mineral nutrition of plants which are of common interest and concern, Mr. D. D. Long, technical adviser in the fertilizer industry, suggested that a symposium be held so that the latest information, both theoretical and practical, associated with the mineral nutrition of plants might be presented by leaders in the different but allied fields concerned. Accordingly, a committee of eleven people representing six national scientific societies and three other agencies dealing with tech- nical matters in this field was organized to plan and arrange such a symposium. Broad representation on the committee gave assurance that the symposium would be national in scope. Sixteen institutions and agencies located in areas from coast to coast, and one each from Hawaii and Sweden, were represented by speakers on the program. More than five hundred persons attended the meetings; ten foreign countries were represented. The Committee expressed a desire to hold the symposium at the University of Wisconsin, because of its central location. This University was particularly happy to accept this proposal; the suggestion came x Preface at the time the University was making plans for celebrating its cen- tennial during the year 1948-1949, and in this connection it was serving as host to a number of symposiums. Fortuitously, this one in Madison was a "natural." In arranging the program, the Committee made an attempt to pro- vide a logical sequence of papers, starting with the soil — the natural source of mineral nutrients — and then treating successively the subjects of entry and translocation of mineral nutrients in plants, the role of minerals in plant nutrition, and finally such modifying influences as light and soil moisture. Two special papers on more practical aspects of the subject were included. Although it was, of course, not possible to treat all phases of this rather broad subject, and many gaps were necessarily left, yet the up-to-date information provided by the con- tributors, all of whom are actively engaged in research, will be of special interest and value to many teachers and investigators, as well as to others concerned with the practical applications. One of the primary objectives of the symposium was to provide an opportunity for presen- tation of latest views regarding the availability of mineral nutrients in soils and mechanisms of absorption and translocation of these mineral nutrients by plants. The papers on these topics should be of special in- terest. The necessary financial support was provided through funds of the Wisconsin Alumni Research Foundation administered by the Graduate School of the University of Wisconsin and by a special grant from the National Fertilizer Association. Grateful acknowledgement is made to these sponsors and to all who took part in the program and served on committees. Especial thanks are due to Drs. L. E. Englebert, Gerald C. GerlofT, and Folke Skoog for invaluable help in reading proof. Madison, Wisconsin EMIL TRUOG September 15, i%i. Table of Contents MINERAL NUTRITION OF PLANTS World Food Possibilities and Fertility Status of Our Soils Charles E. Kellogg and A. C. Orvedal, U. S. De- partment of Agriculture, Beltsville, Maryland PHYSICO-CHEMICAL AND BIOLOGICAL FACTORS AFFECTING NUTRIENT AVAILABILITY IN SOILS Soil as a Medium for Plant Growth 23 Emil Truog, University of Wisconsin The Activities of Cations Held by Soil Colloids and the Chemi- cal Environment of Plant Roots 57 C. Edmund Marshall, University of Missouri The Availability of Soil Anions 79 Roy Overstreet, University of California, and L. A. Dean, U. S. Department of Agriculture, Beltsville, Maryland Contact Phenomena between Adsorbents and Their Signifi- cance in Plant Nutrition 107 Hans Jenny, University of California The Effect of Soil Physical Properties on Nutrient Availability 133 J. B. Page, Ohio State University, and G. B. Bod- man, University of California Role of Soil Microorganisms in Nutrient Availability 167 A. G. Norman, Chemical Corps, Camp Dietrick, Frederick, Maryland 66757 xii Contents MECHANISM OF ENTRY AND TRANSLOCATION OF MINERAL NUTRIENTS IN PLANTS The Nature of the Process of Inorganic Solute Accumulation in Roots 187 T. C. Broyer, University of California The Mechanism of Ion Absorption 251 Hans Burstrom, University of Lund, Sweden The Translocation of Minerals in Plants 261 O. Biddulph, State College of Washington SOME FIELD PROBLEMS IN PLANT NUTRITION Control of Nitrogen Effects on Mcintosh Apple Trees in New York 279 Damon Boynton, Cornell University Production of Vegetable Crops for the Canning Industry 295 Jackson B. Hester, Campbell Soup Company ROLE OF MINERALS IN PLANT NUTRITION Growth and Function as Criteria in Determining the Essential Nature of Inorganic Nutrients 313 Daniel I. Arnon, University of California Mineral Nutrition in Relation to the Ontogeny of Plants 343 W. F. Loehwing, State University of Iowa Correlations between Protein-Carbohydrate Metabolism and Mineral Deficiencies in Plants 3 59 Robert A. Steinberg, U. S. Department of Agricul- ture, Beltsville, Maryland Contents xiii MODIFYING INFLUENCES OF VARIOUS ENVIRON- MENTAL FACTORS UPON MINERAL NUTRITION Light as a Modifying Influence on the Mineral Nutrition of Plants 3 89 Robert B. Withrow, Smithsonian Institution Soil Moisture and the Mineral Nutrition of Plants 411 C. H. Wadleigh and L. A. Richards, U. S. Depart- ment of Agriculture, Riverside, California Environmental Influences on the Growth of Sugar Cane 451 Harry F. Clements, Castle & Cooke, Ltd., University of Hawaii, and Hawaiian Commercial and Sugar Company MINERAL NUTRITION OF PLANTS §b\CAl CHAPTER mm I ^ ^ I World Food Possibilities and Fertility Status of Our Soils CHARLES E. KELLOGG and A. C. ORVEDAL M unkind has lived for a long time on the soils of the world. Despite the examples of spectacular soil depletion that each of us has seen or read about, we still must marvel at the stability of our soils. For many centuries, long before the rise of modern science, eastern and southern Asia had enormous populations. Famines have occurred, to be sure; in fact, this population has rarely been well fed by modern standards, yet it has persisted through the ages. Both western Europe and Japan have maintained increasing populations for three centuries on soils originally low in fertility — soils that have been greatly in- creased in productivity during the last century and a half. What hope, in the light of modern science, can we now give to man that hunger and starvation can be kept from his door? Realizing fully the many critical social, economic, and political obstacles to be over- come (j), what about soils and crop production? Let us look broadly at a few of the physical and biological aspects. ARABLE LAND — AVAILABLE AND POTENTIAL First, do we have enough arable land in the world? If we take a general view of the world's land, we see that about one-half of it is not suitable for cultivation (5). This includes areas covered with everlasting ice and snow, the Tundra, the high mountains, the deserts, and semideserts. There is some significant grazing in the high moun- tains, and water collected in the high mountains is used for irrigating desert lands. In fact, substantial increases in irrigation are possible with 4 Mineral Nutrition of Plants modern methods of soil classification, irrigation, drainage, and fertili- zation. The other one-half of the world's land is only partly arable. Some soils are too stony, too sandy, too hilly, too salty, or too wet for culti- vation. A sharp line cannot be drawn between those which are arable and those which are not. First, any estimate must be based on economic conditions, either consciously or unconsciously assumed. It is physically possible to grow crops almost anywhere: mountainsides can be terraced; stones can be removed; dikes can be built; and water can be carried long distances. Secondly, we do not have a good soil map of the world. Soil maps for several large areas do not exist. In these we can only fall back on informed opinion. Thirdly, the potential use of soil depends upon the associated industrial facilities and transportation. Even estimates of the land now cultivated in the world vary widely around ten per cent. This is because we may start with intensive culti- vation and pass gradually through general farming and extensive farm- ing to nearly wild land with no sharp breaks. Western civilization — our civilization — has grown up mainly in the temperate regions, first in Europe, then in America, and later in Aus- tralia, New Zealand, South Africa, and similar places. It began on the well-watered forested soils near the oceans and great rivers. With the development of technology, it has spread to the interior of continents along the railroads. Now most, although by no means all, of the good land in temperate regions is occupied. We could expand considerably in the United States, probably to a total of around 450 million acres, or possibly even 500 million under conditions of reasonably full employ- ment. This does not count some of the poor soil now in farms that should be used for ranching and forestry. North of the temperate region in the region of Podzol soils, only about one per cent of the land is cultivated. If we increased the per- centage to ten, about 300 million acres of new soil would be available. On the basis of experience in Finland and Scandinavia this seems reasonable, provided transportation and industry are developed along with the agriculture. At least when first cultivated, these soils would not be so fertile as most of those in the temperate region, but experience has demonstrated that they are responsive to management and can be Kellogg and Orvedal 5 well developed for dairying and for vegetables, including potatoes. The great areas of undeveloped soil in the world are in the tropical regions in Africa, South America, Central America, and several of the great tropical islands. In southeastern Asia and India, on some of the Pacific Islands, and in a few parts of other tropical regions, these soils are now intensively used for crops. Yet there are great areas that are hardly touched in relation to their potentialities. It seems reasonable to suggest that at least twenty per cent of the unused tropical soils in the Americas, Africa, and the great islands like New Guinea, Madagascar, and Borneo could be cultivated. This would give us approximately one billion additional potential acres. If we estimated the potential produc- tivity of these tropical soils on the basis of the best results, say in Hawaii and Java, the figure would be almost astronomical. It would be con- servative to use experience in the Philippines as a guide, realizing when we do so that this omits consideration of the great potential increases in efficiency that could come with the application of modern science in the tropics to the same extent we have applied it in the temperate regions. For the northern soils, the 300 million acres north of the temperate region, we may use Finland as a guide. Calculations indicate that with this new land we could more than meet most items listed in the world food needs for i960 by the Food and Agriculture Organization of the United Nations. A few in short supply could easily be increased through some shifts of the agricultural pattern (6). This estimate of 1,300,000,000 acres is probably either too low or too high. A satisfactory estimate cannot be had now. When checked with a reliable soil map, which we may have in a few years, the results will depend a great deal upon the economic assumptions used. Of course, we shall never have an exact figure. At any rate, a large potential acre- age exists. It must be emphasized that these new acres will be "difficult" acres in contrast to some of the land settled by Americans in the past 100 years. There is very little new soil in the world simply waiting for the plow. Most of these new acres will require clearing. They will need careful management from the start. Some of the soils require terraces, some levees, some partial drainage, and some need supplemental irriga- 6 Mineral Nutrition of Plants tion during dry seasons. A large part will need lime, fertilizer, or both from the start. Most of these acres are in the interior of continents, away from regu- lar trade routes or far from good harbors. In order to use them, medical facilities, local industry, and electric power must go along with agricul- tural development. Families settled by themselves or in small groups in the Middle West and in the Great Plains during the nineteenth century. Transportation, industry, and the other services followed them. Very few of these prospective new acres can be settled in this way. On them settlement shall need to be planned with the idea of combined resource development: water, soil, forest, minerals, and power must be considered together. We do not really need all this new land — not yet. During the war careful estimates made by the Department of Agriculture and the land grant colleges showed that it would be entirely practicable to in- crease agricultural production in this country by about twenty per cent on most items and higher than that on several (8). Somewhat lower increases might be expected in some other countries, but in most of them considerably higher ones could be had through the application of what is now known. Thus, even without any new soil the food needs of the world could be met for cereals, roots and tubers, and sugar. But some new areas or further increases in yields beyond those assumed in this earlier study would be needed to supply a bit more fats and oils and considerably more pulses and nuts, fruits and vegetables, meat, and milk. Thus, taking the two together — the potential new soil and the dem- onstrated potential production under good management on land al- ready being farmed — the world could have food far beyond the amounts estimated as required for the world population in i960. Now, such estimates may appear to be very optimistic. They prob- ably are in terms of the real political and economic situation that we see about us. They indicate what could be done with our present knowl- edge // adequate institutional arrangements were made for effective soil use on a sustained production basis. In another sense they are low. They are low because these estimates take no account of entirely new tech- nology. They assume merely a general acceptance of existing technology Kellogg and Orvedal 7 as now used by millions of farmers. Yet we know that the efficiency of farm production is increasing at an accelerated rate. It has increased considerably since the estimates were made by the Department of Agri- culture and the land grant colleges. And we must recall again that the level of production we are seeking will not be based upon history, nor based upon any mythical "natural balance." // is that level of production on a sustained basis made possible by modern science and technology in a peaceful world with reasonably full employment. We seek an effective cultural balance between people and resources. We in the United States and in western Europe have been inclined to take for granted the enormous increases in efficiency of production during the past 150 years. These have been reflected in yields for a long time in Europe although not in the United States until recently. There are several reasons for this difference. In the United States land has been relatively plentiful and labor relatively expensive, except during periods of depression. Alert farm managers are concerned as much with reducing inputs as with increasing outputs. Many of our most important improvements are designed to reduce labor even at some sacrifice of total harvest. During the latter part of the nineteenth cen- tury and the early part of the twentieth century, many millions of acres of new soil were brought into use in the subhumid and semiarid re- gions where normal yields are relatively low. Then, too, many of our plant breeding programs have been concerned with widening the range of soil types on which important crops may be grown rather than simply increasing the yields on a few soil types. As a result, our farm- ers have many more choices of crops than they had formerly. Some farmers have allowed their soils to deteriorate seriously, even while suc- cessful farmers were improving theirs.* Yet even so, there have been * Erosion, decline of organic matter, increasing acidity, loss of soil structure, increasing salinity, soil blowing, and especially declining plant nutrients, are all in the picture. However, there has also been improvement on many farms through the use of lime on soils naturally acid, the use of fertilizers on soils naturally low in plant nutrients, drainage, irrigation, the introduction of legume meadows, and livestock farming. No one knows what the net result has been. Nor would it make much difference to these estimates if we did know, since our efficiency is so much higher than it was 200 years ago and so much lower than it could be 8 Mineral Nutrition of Plants significant increases in the average yields of several of our major crops in the last few years, aside from climatic effects. As applied to the United States, these effects of technology are much more clear when considered in terms of efficiency (2). For example, between 1800 and 1940 the number of man hours neede.d to produce 100 bushels of wheat dropped from 373 to 47. Similar improvements were made with the other major crops. In 1820, after science had al- ready had some effect, one farm worker supported about four and one- half other people. In 1946 the figure was fourteen and one-half. Modern science has already increased our efficiency and continues to do so at an accelerated rate. This has happened in the temperate re- gions. There is every reason to believe that similar potentialities for the application of science to agriculture exist in the far north and especially in the tropics. The tropics have some handicaps but they also have many advantages. Our guess would be that science will be even more influ- ential in the tropics than it has been in the temperate regions. The great need there is for research institutes devoted to fundamental investiga- tions. The soils are so different from those in the temperate regions that technology can be transferred only to a very limited extent. SOIL FERTILITY OF THE GREAT SOIL GROUPS Following this general view of the potentialities for food production, let us look more specifically at soil fertility. Plant nutrients tend to be- come limiting wherever man uses soils for crops, but the fertility prob- lems vary in both kind and intensity from place to place. Different crops have unlike nutrient requirements, and different soil types vary in their capacity to provide nutrients. Since many thousands of unique soil types exist in the world, exceptions must be permitted in any gen- eralization. Yet, a broad view of the fertility status of the principal great groups of soils may help us measure the fertility problems in world food production. with what we know now. Of course, we do need to learn these relationships specifically, farm by farm, to get on with the job of preparing recommendations for sustained production. Kellogg and Orvedal 9 Gray-Brown Podzolic soils The Gray-Brown Podzolic soils are dominant in the agriculture of the United States from Minnesota east and Tennessee north, and in the northwestern part of Europe. Like other podzolic soils, these are leached, acid, and relatively low in most plant nutrients in available form. They are also rather low in organic matter. Despite these fertility deficiencies, they are highly responsive to management. With lime and fertilizers, they support a very wide range of crops. We have learned to make from them fertile, arable soils that give high yields. Phosphatic fertilizers and lime are needed on nearly all of these soils. The amounts of potassium and nitrogen required depend upon the legumes grown and the animal manures available; but, usually, at least some chemical fertilizers containing these nutrients are required for optimum efficiency. In many places magnesium and boron are needed for some crops. Responses have been reported for most of the nutrients, but, generally speaking, deficiencies of the minor elements are less common than on the soils of warmer regions. Western civilization grew up on the Gray-Brown Podzolic soils and expanded from them onto others. Most of the research on soils and soil- plant relationships, especially during the nineteenth century, was con- ducted on these soils and their close relatives. The results have been dramatic. From the fall of Rome nearly to the French Revolution, grain yields in Europe were around six to ten bushels to the acre. With the adoption of crop rotations, they nearly doubled. Not only that, the elimination of the fallow year gave more acres for harvest. The appli- cation of chalk and farm manures became general. In Germany, wheat yields went to about 16 bushels to the acre by 1850. At the same time, they were 14 bushels to the acre in France and somewhat over 20 bushels in Britain. By 1906, they had gone to 30 bushels in Germany and to over 30 in Britain, while they were about 20 in France. Now they stand at about 35 in Britain (4). Something more than one-half of the increases in Britain and Germany came before the common use of chemical fertilizers after 1850. This remarkable increase in production has been attained through the use of improved techniques, most of which have grown out of io Mineral Nutrition of Plants scientific research over the last 150 years. On no other great soil group has a comparable amount of research been clone and on no other can we find similar, widespread increases in production, although there are important advances in other areas that point to the potentialities. Scientists and farmers alike, on the Gray-Brown Podzolic soils of Europe and the United States, know how to overcome the handicaps of relatively low fertility by liming, fertilization, growing legumes, and so on. The question of what yields to strive for is, within limits, an economic matter of input-output relationships of the various factors involved, including fertilizers. Yet, even now, only a part of the farm- ers on these soils are really efficient. If all of them followed the prac- tices of the most efficient one-quarter or one-third, production could probably be increased another 25 to 35 per cent, to say nothing of possible contributions of entirely new technology. Ch ernozems The Chernozems are the great wheat-producing black soils of the United States, Canada, Argentina, and the. Soviet Union. Unlike the Gray-Brown Podzolic soils, the Chernozems are highly fertile at the start. They have abundant organic matter and an excellent granular structure. They are rich in nitrogen, calcium, potassium, and most other plant nutrients. In spite of this storehouse of nutrients, yields on Chernozems are not especially high, on the average. Rainfall on these soils is erratic and often low. So far, science has had conspicuously less success in over- coming the limitations set by drought than those set by low fertility. Still, a lot has been done recently to increase yields on Chernozems, especially through improvement in varieties of grains and through better and more timely tillage with powerful machines. Yet, there are fertility problems with Chernozems. Under the pre- vailing practice of continuous grain production, the nitrogen supply declines. The growing of legumes, like alfalfa and sweet clover, tends to offset this loss, but it also tends to deplete the moisture supply for the following grain crop. Thus, yields following legumes are sometimes lower than they are following the previous grain crop and especially following fallow. No doubt, we shall see more nitrogen fertilizer used Kellogg and Orvedal n on these soils, but chemical nitrogen will not be enough. The organic matter must also be maintained through the proper use of crop residues. Years ago, poor results from experiments with rock phosphate on Chernozems led some to conclude that they would not respond to phos- phate fertilization. Chernozem soils are too rich in calcium and too high in pH for rock phosphate to become available. But later results indicate that strong feeders, like sweet clover and alfalfa, may show some responses even from rock phosphate. Greater increases, however, are obtained from the more soluble phosphates, especially for small grains and corn. This is partly because of depletion caused by long, continuous cropping, but more because of better varieties and better tillage practices that have made it possible for plants to use more phos- phate efficiently. The use of phosphate is increasing and should increase a great deal more. The need, however, is likely to vary a good deal from one local type to another. Potassium generally is not needed on Chernozems, although there are exceptions. It cannot be ruled out. Deficiencies of the minor nutrients are uncommon but may be expected here and there, if other fertilizers are used abundantly and especially with supplemental irrigation. Desert soils The soils of the desert are well supplied with mineral plant nu- trients, except for the very sandy ones, but are low in organic matter and nitrogen. Often the mineral nutrients are poorly balanced, with excesses of some. Without irrigation, these soils are used only for ex- tensive grazing, and moisture is the limiting factor for plant growth rather than fertility. Where irrigated, however, fertility problems be- come highly important. With technically sound irrigation and opti- mum moisture conditions, the ceiling set on nonirrigated soils by the moisture supply is lifted and other factors become limiting. Of first importance is nitrogen. With legumes in the rotation, some nitrogen may be supplied by them, but ordinarily commercial fertilizer is also needed. Potassium is usually abundant in desert soils, but phos- phorus is commonly low. Then too, deficiencies of the minor nutrients, like iron and zinc, are common. These conditions are partly stimulated by the high amounts of lime generally present. The fertility problems 12 Mineral Nutrition of Plants of irrigated soils are often further complicated by excesses of soluble salts, of salts in general, or of some specific toxic salt like borax. Here is an important field where soil science, plant nutrition, and plant breeding come together. A good deal of active research is under- way but more is needed. In view of the high yields that can be obtained with improvements in combinations of practices, we can expect the Desert soils to make significantly larger contributions to the world food supply. Latosols * Vast areas of potentially arable soils exist in the tropics, and here may rest a great deal of the future hope of mankind for abundant food; but the problems are great, and the research to guide us is little. These red soils of the tropics, Latosols, are strongly weathered and highly leached. By standards used in temperate regions, they are low in all plant nutrients. Furthermore, they have low base-exchange capac- ities and high phosphate-fixing capacities. The tropics, nevertheless, have several advantages. There is little or no frost. The growing season is as long as the moist season, say six, nine, or even twelve months. The limitations imposed by climate, there- fore, are generally less severe than in the temperate and cold regions. Already, we have indications of some peculiarities of management requirements and a hopeful glimpse here and there of production attainable under improved management. Important in connection with fertility is the mineral cycle — the up- take of minerals and their return to the soil. Plants in the tropics grow rapidly and decay quickly. If an ion of calcium, let us say, were to cir- culate three times in the tropics to once in the temperate regions; the same ion would theoretically do three times as much work in one place as the other. Equal amounts of plant nutrients, therefore, might go farther in the tropics than in temperate regions. Since the Latosols of the tropics have low base-exchange capacities and, hence, low nutrient-holding capacities, a continuous cover of plants *A term recently introduced to apply to a broad group of soils in the tropics that are red, leached, relatively high in iron and aluminum as compared with silica, relatively low in base-exchange capacity, and highly aggregated. Kellogg and Orvedal 13 is almost a necessity. Under such cover, the tendency of rapid leaching of nutrients will be offset greatly by the equally rapid uptake of these nutrients by plants. Thus, the absorbed nutrients will again be brought to the surface and subsequently released upon the death and decom- position of the vegetative material. Generally, the common practice in the temperate regions of growing a single crop in a field, with periods without plant growth between harvest and sowing, is unsuited to the humid tropics. We need there to think of mixed cultures rather than monocultures, even rather than rotations of crops grown in mono- culture. Exposure of the soil to direct sunlight has been observed to be harm- ful to many tropical soils. Nearly continuous shade is essential for the maintenance of productivity. This is partly a matter of fertility and partly a matter of soil structure. The removal of all vegetation in some places even brings about severe drying at and near the surface, resulting in formation of a hard laterite crust or pan in soils such as Ground- Water Laterites. The corridor system of crop rotation, perhaps best developed in the Belgian Congo, shows great promise as a good system for many tropi- cal soils, especially where commercial fertilizers cannot be had eco- nomically. This system is simply a modification of the old practice of shifting cultivation. Under the corridor system, a given tract of arable land is divided into long strips, say 1,000 yards long and 100 to 300 yards wide. A corridor is provided for each year of cropping and for each year under forest fallow. These are arranged so that some stage of forest borders each corridor in crops. Besides, a few specimen trees may be left for partial shade and regeneration. On a comparatively good soil, one may have a six-year rotation of crops with twelve years of forest fallow. Such an area needs eighteen corridors. Each year an old corridor is returned to forest fallow and a new one cut and planted to crops (9). Good results have been obtained from organic manures and com- posts in India (7) and elsewhere. Although the use of manures has increased yields in the temperate regions too, they are of special im- portance in the tropics because of the gradual release of nutrients in the organic matter as contrasted to mineral fertilizers. Where leaching . A R Y *4 Mineral Nutrition of Plants n Prairie soils, Degraded Chernozems Dark Groy and Black Soils of Tropical Sovannos (with some inclusions of Chernozems and Reddish Chestnut Soils) Sierozems, Desert, ond Red Desert Soils Chernozems and Reddish Chestnut soils (with somemctus ions of dork gray and black soils of tropicol sovannos) Chestnut, Brown, and . Reddish Brown Soils Podzols (with much bog ) Figure i. Provisional schematic soil map. Kellogg and Orvedal ^ Groy - Brown Podzolic Soils, v/sy///tf\ Brown Forest Soils, etc -j Soils of Mountains and - '- :..>v;1 Mountain Volleys (complex ) Alluvial Soils (many smoll but important areas, not shown on mop, occur in all parts of the world ) LotosolsfRed Lateritic. Reddish Drown Lateritic, etc.) Red Yellow Podzolic, Terro Rossa, etc. Tu n d r o 1 6 Mineral Nutrition of Plants is severe and exchange capacities low, this difference may be quite important. Further, the balance among nutrients, important everywhere, is espe- cially so in tropical soils. Much more research is needed to establish the correct fertilizer ratios to use on the many thousands of local soil types. The level of all nutrients is so low, in absolute terms, that an excess of one, caused by liming or fertilization, can easily upset the balance. Thus, in our ignorance, the use of compost has an advantage. Where a mixture of normal plants is used, one is bound to get a reasonably balanced supply of plant nutrients with decomposition. Someday we shall know from research the proper chemical mixtures to use. Yet, these mixtures can never substitute entirely for the good effects of mulch and the effects of organic matter in maintaining soil structure. The large yields obtained with heavy fertilization on tropical soils in Hawaii, northern Queensland, and other places, where the products of industry are available in the tropics, give us a glimpse of enormous potentialities, provided agriculture and industry are developed together. In the meantime, much can be done in the tropics to improve fertility and increase production without the use of mineral fertilizers. Our guess would be that scientific agriculture will increase yields even more in the tropics than it has on the Gray-Brown Podzolic soils of western Europe, and only a part of this will be due to fertilizers. Besides the four major groups of soils discussed here, there are many others, each with unique fertility problems. But most of these soils are intergrades among the four broad groups. SOIL DISTRIBUTION AND FERTILITY PROBLEMS The provisional soil map shown in the accompanying figure gives a broad view of how the various kinds of fertility problems are distri- buted. This map is schematic and highly generalized; it is not detailed enough for local predictions of land-use potentialities. Gray-Brown Podzolic soils and their associates are shown in the northeastern part of the United States, north-central Europe, and China. In relation to the total land area of the world, the extent of these soils is small. Yet, on them, both food production and fertilizer consumption are high, especially in Europe. In the prewar years, Kellogg and Orvedal 17 Europe alone used 59 per cent of all nitrogen, 63 per cent of all phos- phoric acid, and 78 per cent of all potash in the world's fertilizer. Within Europe, these fertilizers were used largely in Germany, France, Denmark, Britain, and the Low Countries (/). North of the Gray-Brown Podzolic soils is a great area of Podzols and their associates in North America and Eurasia. Large percentages of these are too stony, too sandy, too hilly, or too swampy for feasible cultivation. A large part of the area has only a very short growing season. However, some of the Podzols are under cultivation, and more could be cleared and cultivated. The fertility problems are similar in kind, but perhaps greater in intensity, to those of the Gray-Brown Pod- zolic soils. At least, lime and fertilizers are usually needed almost at once after clearing, for efficient production. Between the Gray-Brown Podzolic soils and the Chernozems are the highly productive Prairie soils and Degraded Chernozems. They are less leached than the Podzolic soils but have more moisture than the Cherno- zems. The fertility problems are also intermediate. After some period of cultivation, the length of which varies with the local soil type, lime or fertilizers or both are required for efficient production on many of the local soil types. Phosphates are of first importance, generally, in mixed farming. Even then, supplemental nitrogen often gives responses on corn. The Chernozem and Reddish Chestnut soils are found mainly in Eurasia, North America, and South America, although there are im- portant areas in Australia and elsewhere. These are the great wheat- producing areas. Besides the Chernozems, there are Dark Gray and Brown soils in the tropics. These are also fairly rich in mineral nutrients, more so than the Latosols, but they are relatively low in organic matter and nitrogen despite the dark color. On the dry side of the Chernozems are the Chestnut, Brown, and Reddish-Brown soils. The fertility problems here are intermediate be- tween those of Chernozems and those of Desert soils. Vast areas of soils in deserts and semideserts are found in Africa, Asia, and Australia. Except for very extensive grazing, agricultural use is limited largely to the irrigated places. Although these are very im- 18 Mineral Nutrition of Plants portant in the aggregate, they are too small to be shown on a map of this scale. Finally, we emphasize the vast areas of Latosols, especially in Africa and South America. These are intensively used for crops in southeastern Asia and in nearby islands, in Hawaii, in parts of northern Queensland, and in many other small areas in Central America, South America, and Africa. Nevertheless, great regions are little developed. Even if crop production were increased on the Latosols only as much as it has on the Gray-Brown Podzolic soils during the past 150 years, the results would be enormous. The job is difficult; otherwise, it would not have waited so long. LAND AND FERTILIZER POTENTIALS It is obvious that any substantial increase in food production by boost- ing yields on old land, and by bringing in new, will depend upon vastly increased fertilizer use. What about potential fertilizer materials? Two years ago, Salter (6) made some estimates along this line. These included not only 1,940,000,000 acres of existing crop land, but also the suggested 1,300,000,000 acres of new land. Nitrogen can be obtained from the air wherever nitrogen-fixing plants can be built. Thus, for nitrogen, industrial facilities, not raw ma- terials, are limiting. Known reserves of mineral phosphate and potash were estimated to be adequate for 2,000 and 500 years, respectively, within an assumed rate of phosphate use eight times the present one, and of potash use, eighteen times. These estimates, of course, take no notice of probable deposits in the many inadequately explored parts of the world. Nor do they allow for increasing efficiency in the use of fertilizer, which also can be expected. Although fertilizers are in short supply, the raw materials for them are abundant. CONCLUSIONS In our view, the soil problems, including the problems of fertility, are manageable in the biological and physical sense. Through the ap- plication of science and the expansion of research where needed, no predictable limit of production can be foreseen. Of course, there must be one, but it is very high. Kellogg and Orvedal 19 This is not saying, however, that mankind will be well fed. The social, economic, and political problems are many and difficult. The technical problems of soils, plants, and animals, great as they are, are small by comparison. Perhaps the really big question is: How badly do we want abundant food in the world ? How much are we willing to sacrifice now, as individuals, as groups, and as nations? These are the sorts of questions that must be debated in terms of value judgments. They cannot be answered scientifically. What soil science says is that if people want an efficient agriculture, producing abundant food on a sustained basis, and are willing to develop the necessary social institutions, they may have it. REFERENCES 1. Clark, K. G., and Sherman, Mildred S., "Prewar World Production and Consumption of Plant Foods in Fertilizers," U . S. Dept. Agr. Misc. Publ. $gj (1946). 2. Cooper, M. R., Barton, C. T., and Brodell, A. P., "Progress of Farm Mechanization," U. S. Dept. Agr. Misc. Publ. 630 (1947). 3. Kellogg, Charles E., "Food, Soil, and People," UNESCO Food and People Series 6 (New York, 1950), 64 pp. 4. Ogg, W. G., Fertiliser, Feeding Stuffs and Farm Supplies /., 34:329 (1948). 5. Prassolov, L. I., Pedology (U.S.S.R.), 2:69 (1946). 6. Salter, Robert M., Science, 105:533 (1947). 7. Stewart, Alexander B., Soil Fertility Investigations in India with Special Reference to Manuring (Delhi, Army Press, 1947). 8. "Peacetime Adjustments in Farming," U. S. Dept. Agr. Misc. Publ. 595 (i945)- 9. "Lands of Shifting Cultivation in Soil Conservation: An International Study," Food and Agr. Organization of the United Nations, Agr. Ser., 4:110 (1948). PHYSICO-CHEMICAL AND BIOLOGICAL FACTORS AFFECTING NUTRIENT AVAILABILITY IN SOILS CHAPTER /- Soil as a Medium for Plant Growth EMIL TRUOG o ur best soils provide a well-nigh perfect medium for the growth of crop plants. Why should this be the case? Were soils specifically designed and created to support plants of the type we now have? More likely the correct answer is that plants by evolution gradu- ally adapted themselves to grow on the soils as they happened to exist. This is substantiated by the belief that the first or primitive plants lived in water from whence one or more species migrated to the land, and there, by evolution, gradually developed into the many forms of land plants which today, as regards number of species, far surpass those of the ocean. Whatever the answers may be to the questions just raised, the fact that nearly all our food and clothing and much of our housing come directly or indirectly from the soil provides ample reason and incentive for learning why some soils are good and many are poor media for crop growth, and how the poor soils may be made better and all good soils conserved indefinitely. Of the natural resources which are easily subject to serious deterioration and even complete destruction, soil is the most precious of all. To be sure, sunshine and water are just as important or even more so than soil, but fortunately neither is subject to destruc- tion by the carelessness of man, although inland water resources may be greatly impaired through improper soil management practices. The soil may, of course, be studied from several standpoints. The pedologist thinks of soils primarily in terms of their origin, form or morphology (profile characteristics), and classification. The soil physi- cist, the soil chemist, the soil microbiologist, and the soil conservationist each has his own field of special interest and study. A well-rounded 24 Mineral Nutrition of Plants soil scientist should have at least a working familiarity and knowledge in all of these special fields. On the other hand, many agronomists and horticulturists are interested in soil science solely from the standpoint of knowing how soils may serve as a medium for plant growth, and how they must be managed so as to produce satisfactory crops and at the same time be conserved for all time for this purpose. Although this paper is devoted primarily to matters of special interest to the agron- omist, horticulturist, and plant physiologist, a brief discussion of soils from the pedologic viewpoint seems appropriate. SOIL AS VIEWED BY THE PEDOLOGIST The so-called science of pedology is the basic science of soils which deals with the origin, evolution, morphology, and classification of soils without any reference necessarily to their use for crop production. Soils are considered to be natural bodies or entities in themselves, deserving study as such, just as plants are to the botanist who often studies them without thinking of their economic value or use. This does not mean that the pedologist in pursuit of his vocation does not add to the eco- nomic welfare of society. Quite the contrary. By learning as much as possible about soils just as soils, a sound basic foundation is provided for their best use in the same way that the botanist provides this in the case of plants. In order to determine the main characteristics of a soil so that it may be properly classified and evaluated, more than surface examination is needed; it is, in fact, necessary to expose the soil for observation and study from top to bottom. The pedologist calls this exposure a soil profile, which is simply a vertical cross-sectional view of a soil, extend- ing from the surface to a depth, usually, of three or four feet, or to bed- rock. It is the view one may see along a road cut, such as that pictured in Figure i, where it will be noted in particular that the soil has three layers, called horizons by the pedologist. It may well be asked — Why does a soil have layers or horizons which differ in characteristics? Why is it not uniform throughout? A consideration of how a soil forms and develops will suggest answers. Let us suppose that an area of bare rock has become exposed to the action of air and water with accompanying ever-changing temperatures Emit Truog 25 in a climate like that of Wisconsin. Under this action, cracks, fissures, and flaws develop in all rocks, into which seeds of plants carried by wind and birds may lodge, germinate, and grow. All of us have seen examples of this in a stone quarry, stone wall, or right in our rock garden, and have marveled at the ability of certain plants to anchor themselves in minute cracks and then flourish on virtually bare rock. Lichens and moss, which will grow on practically bare rock, are often the forerunners of the higher plants. The dense and compact growth of moss provides favorable conditions for the accumulation of both mineral and organic matter. Having gained a foothold, the plants help to hold the rock powder being formed in place against removal by wind and water, and in addition exert a powerful influence in accelerating the further disintegration of the rock. Once established, the plants start to shed and cast off leaves, roots, and other parts which become mixed with the rock powder and provide organic matter which serves as food for microorganisms brought in by wind, water, and other agencies. Thus the embryonic soil receives the touch of life which enables it to grow to maturity. However, even after the beginning just described has been made, it may take 1,000 years to produce an inch of soil material. After a layer, usually eight to twelve inches deep, has developed, the further produc- tion of soil material by weathering of rock material produces a second layer which is quite different from the one above; it receives very little organic matter, being somewhat distant from the main source of such material, and its population of microorganisms is thus much lower; also, it is less subject to leaching and may, in fact, receive fine clay and other material from above. In time this layer becomes as thick as or thicker than the layer above, but, because of a lack of organic matter which gives all or a portion of the layer above a dark color, it usually has a yellowish, brownish, or mottled brownish to bluish color, de- pending upon the state of oxidation of the iron as determined by drainage and aeration. Of course, the bedrock beneath keeps on disintegrating under the solvent action of water and the action of frost and other agencies. The weathering rock material exists first largely as a layer of loose gravel which receives additions continually from below as the bedrock dis- 26 Mineral Nutrition of Plants integrates, but also loses material as its upper portion weathers more completely from gravel to soil and becomes incorporated as a part of the layer above. Thus we see that in the development of soils from rock, it is natural and inevitable that the characteristics of the soil should vary from the A- HORIZON TOPSOIL B- HORIZON SUBSOIL C- HORIZON PARENT MATERIAL BEDROCK Figure i. A vertical cross-sectional view of a soil showing the horizons (layers) that make up a soil. This is called a soil profile. surface downward. Although this variation is gradual, it is sufficiently marked in most cases to allow ready observation of three layers of the kind described. For convenience of reference as illustrated in Figure i, the pedologist calls the upper layer the A horizon, the middle layer the B horizon, and the lower layer just above the bedrock the C horizon. Strictly speaking, the true soil, called the solum, consists of horizons A and R; horizon C is raw material out of which the soil proper is made. Emil Truog 27 CHARACTER OF SOILS DETERMINED BY FIVE FACTORS As is to be expected, great variations exist in the characteristics of the respective horizons of soils found in various places. These varia- tions are due to the extent to which the following five main factors involved in soil formation have come into play. These five factors are climate, native vegetation, parent rock material, topography or slope, and age or maturity of the soil. The first two — climate and vegetation — are active factors and have to do with dynamic forces, such as weather- ing and leaching. The other three factors are of a passive nature and do not represent forces or activity in themselves, but they influence greatly the effects of the active factors on the product which remains and, hence, the character of a soil as found and its suitability as a medium for crop growth. From a broad or world viewpoint, climate, which not only determines the rate of weathering and leaching but in turn also largely the kind of vegetation which grows, has the greatest influence on the general type of soils formed, particularly as regards the organic matter content and the color. It determines which of the so-called zonal soils will be formed, namely, podzol, chernozem, or laterite. However, within a local area, such as the northern half of Wisconsin which lies within a single climatic zone, parent rock material is the determining factor for the texture of the soils formed. In a cool moist climate like that of northern Wisconsin where leach- ing is severe and coniferous trees predominate, the podzols or highly podzolized soils predominate. Whether of a sandy or heavier texture, they are characterized by an ash gray A.2 horizon and a brownish, yel- lowish, or yellow-blue mottled B horizon, which usually has a heavier texture than the A horizon. These soils, having been drastically leached, are strongly acid and low in available plant nutrients, particularly potassium. As media for crop growth, they are generally not satisfac- tory until limed and heavily fertilized with potash. The subsoil or B horizon of the heavier soils is often inadequately drained and aerated for many crops. Compared with the northern part, the climate of the southern part of Wisconsin is quite different. Here, higher temperatures have caused 28 Mineral Nutrition of Plants more evaporation and transpiration and, hence, less leaching. Also, the higher temperatures have favored hardwoods rather than conifers, and this has further reduced leaching effects. As a result, the soils, being less leached, are less podzolized and, therefore, better media for crop growth. The fact that the parent rock material in the northern part of the state has been sandstone and igneous rocks and in the southern part largely limestone has also greatly favored the formation of better soils in the latter area. If one goes south to humid semitropical or tropical regions, reddish or lateritic soils are quite generally encountered. These soils are usually low in organic matter and have generally suffered severe leaching. They are, however, generally well drained and aerated and when properly fertilized usually serve as excellent media for crop growth. Going west from Wisconsin, one encounters regions of less and less precipitation. On reaching the subhumid region of the Dakotas where the annual precipitation averages 20 to 25 inches, it is found that tall grass vegetation has flourished instead of forest. Under these condi- tions, the chernozem and related soils were formed. They are rich in organic matter, have suffered little or no leaching, are rich in plant nutrients, and in many respects represent our best media for crop growth. The main failing is that the precipitation which accompanies them is not at all dependable. Thus we see that the climate which promotes the formation of the best soil media for crop growth usually does not have sufficient precipitation for highest crop yields. To the west of the subhumid region of the Dakotas, one encounters the semiarid region where the average annual precipitation is less than 20 inches. Here, the limited soil moisture supply has restricted vegetation to the extent that the soils formed are practically devoid of organic matter. However, because of the virtual absence of leach- ing, these soils are usually well supplied with available nutrients and often make good media for crop growth when irrigated. Because of little or no leaching, these soils are generally alkaline, frequently to the extent of injury to crops. (See references 3, 4, 8, and 9 for further details regarding origin and classification of soils and other pedologic matters.) Emil Truog 29 THE CONSTITUTION OF SOILS Turning now from the brief consideration of the origin, gross morphology, and kinds of soils, there follows a discussion of the chemi- cal, physical, and mineralogical constitution— that is, what might be called the anatomy and histology of soils. Matter exists in three physi- cally distinct conditions or states commonly referred to as the solid, liquid, and gaseous phases. It is important to recognize that all three phases of matter, and in addition a "living phase" consisting of soil bacteria and other organisms, enter into the constitution of a soil when it is in proper condition to function as a medium for plant growth. Thus, in the language of the physical chemist, soils are three- phase systems: there is the solid phase consisting of innumerable minerals and organic substances; the liquid phase consisting of the soil moisture or water in which relatively small amounts of the solid phase dissolve; and the gaseous phase, the soil air, which fills the pore space not occupied by the soil moisture. The total volume of the latter two in any soil must, of course, always equal the volume of the soil's pore space, although the volume of each fluctuates with the moisture content of the soil. But a soil serving as a medium for plant growth is more than a three-phase system in the ordinary sense: it teems with microorgan- isms, and has, in addition, what may be called "a living phase," which must always be taken into account in any consideration of the soil as a medium for crop production. The constitution of the three con- ventional phases and also the "living phase" of the surface or plow layer of a typical silt loam soil is given in outline form in Table I. The solid phase It will be noted that the solid portion or phase of a silt loam oc- cupies only about one-half of the volume of the soil, the remainder being, of course, pore space when the soil is free of moisture. The pro- portion of solid-to-pore space varies greatly among the different types of soils. In general, the heavier the soil— that is, the greater the content of finer material, especially clay, and also the greater the content of organic matter— the higher will be the percentage of pore space. Thus, TABLE i Constitution of the Soil (Plow Layer) of a Typical Cultivated Silt Loam at or near Optimum Moisture Content I. Solid Phase: The soil proper, 50 per cent by volume Inorganic portion, 95 per cent by weight Coarser portion: Sand, Fine clay or colloi- silt, and coarse clay, par- ticles >o.2(jl diam., 80 per cent by weight. Largely primary miner- als: Feldspars, micas, quartz, and many others dal portion: Particles Tr»rf> /it ^P7^ r»f racpc nppn to he al-isnrhpn as 190.90 carbonate and bicarbonate. Timothy f A1203 0.053 3.12 Si02 1. 851 61 .64 Fe203 0.155 5.82 CI 0.583 16.44 CaO °-332 11.84 p2o5 0.474 20.03 MgO 0.220 10.91 so3 0.460 1 1 .50 K20 2. 190 46.50 N 0.990 70.71 Na20 0.017 0.55 MnO 0.006 0. 11 Total 78.85 Total 180.32 ^— = 0.563; therefore, 56.3% of acidic constituents may be 0,32 absorbed as free acid and none of basic constitu- ents need to be absorbed as carbonates and bicarbonates. *Data for alfalfa are based on analyses by O. C. Magistad of five Wisconsin samples cut at hay stage. fThe figure for nitrogen content of timothy was taken from Henry and Morrison, "Feeds and Feeding." All other data for timothy are based on analyses of three field samples cut in bloom stage and reported in U.S.D.A. Bur. Soils Bui. 600 (1917) p. 11. Emil Truog 39 An examination of the analytical data for timothy reveals a much different situation in this connection. This plant (on a milliequivalent basis) requires more than double the amount of the acidic than of the basic constituents. There may be some question as to the propriety of including silica (Si02) and chlorine in these calculations because of the weakness of the former as an acid and nonessentiality of both as plant constituents. However, even when these two constituents are disre- garded in the calculations, a considerably greater equivalent amount of acidic than of basic constituents is still required by timothy. Thus, the timothy plant can absorb a considerable portion of its nitrogen and sulfur needs in the form of free nitric and sulfuric acids without up- setting the internal nutrient balance of needed acids and bases. In the above considerations it is assumed that all the nitrogen derived from the soil is absorbed as the nitrate. Absorption by plants of some of the nitrogen in the ammonium form as a cation will not alter ma- terially the deductions made, since alfalfa will still have a much greater base requirement than timothy. It has been noted that alfalfa often grows much more satisfactorily on distinctly acid soils if these soils happen to be well supplied with available nitrogen. However, in the case of nonacid and very slightly acid soils, the supply of available soil nitrogen is not crucial providing the alfalfa is properly inoculated. The foregoing discussion accords with these observations. When the alfalfa has access to soil nitrogen, much of the nitrogen absorbed carries a base with it even though the soil is acid. However, if the plant is dependent upon atmospheric nitrogen, then no bases are carried in with the nitrogen, and the plant must compensate for this by absorbing from the soil considerable amounts of the required bases in the form of the carbonate or bicarbonate, such as calcium bicarbonate. The carbonates and bicarbonates, because of the weakness of the acid radical, function in the plant much like a free base. Only when soils are not very acid can they be obtained in ade- quate amounts for plants like alfalfa which have a high requirement. The indirect influence The indirect influence of soil reaction on plant growth may be exerted in a number of ways. Of tremendous importance is the inflq- 40 Mineral Nutrition of Plants ence due to its effects on the availability of the plant nutrients. Strong acidity is, in general, unfavorable, while very slight acidity (about pH 6.5) is favorable. Figure 3 presents the subject diagramatically. It will be noted that at strong acidity, there is a marked drop in the supplies of available nutrients, which may exist as exchangeable cations. This is, of course, to be expected, because with increasing acidity, increasing amounts of these nutrient elements which might exist as exchangeable cations become replaced by hydrogen. In certain ranges on the alkaline side, some of the nutrient elements become much less soluble and avail- able. A full discussion of this subject is beyond the scope of this paper and may be found elsewhere (//, 12). The indirect influence of soil reaction on plant growth because of its effects on the activity of soil microorganisms is extremely important, particularly as related to the availability and fixation of nitrogen. In general, the more desirable types of biological activity in soils are pro- moted by a soil reaction that is neutral or nearly so. Increasing acidity favors the solubility of aluminum and manganese, and also copper and zinc and other heavy metals should they happen to be present in undue amounts. When soil acidity becomes more in- tense than pH 5.0, concentrations of soluble aluminum and manganese which are toxic to certain plants may occur in some soils. Unless built up through the extensive use of spray materials or the addition in other ways, toxic concentrations of copper and other heavy metals seldom if ever occur. The physical condition of heavy soils, especially, is generally known to be affected unfavorably by strong acidity which gives rise to a cal- cium bicarbonate supply in the soil solution which is insufficient to keep the clay well flocculated — a condition that is necessary for full promo- tion of the desirable granular or crumb structure. In sandy soils, a high level of available lime (low acidity or absence of it) may help to coun- teract excessive looseness by acting as a binding agent. Improvement of the physical condition of soils in the ways indicated helps to promote a satisfactory regulation of the air and moisture supply of soils; this, in turn, promotes the more desirable types of biological activity and chem- ical reactions. A highly acid or alkaline condition, by inducing defloc- culation, causes a movement of colloidal material from the surface soil ~Em.il Truog 41 - £ Ji "« « ^ 3 o as Oh >^ 3 u C u rt rt a. c o o .2 c r3 -a o -r -a c « V) > ■4-f 3 hn a! -a c 4J O W5 o ID s "Sh u 03 r* a, o ^"^ '-4— 1 W !_) Oh -a c 03 V 03 Hi a o 03 2 o 03 u 3 — >. O 04 wT Ih TD -o M-l n 4J c flj 4— > u 03 'o c oj o J-1 o u. 4-» O c .2 « 3 c C 03 03 -a V u c V O u, 3 "> -C 03 3 _C 5 3 u U (J 03 U r3 Oh 03 > ^ « r3 >^ o3 73 3 "O ™ 03 u V V 3 G ^ •- 3 G oj 42 Mineral Nutrition of Plants into the subsoil where it may be precipitated as a hardpan. The deple- tion of the surface soil of its colloidal material and the formation of a hardpan have an unfavorable influence on soil fertility. Some soil-borne plant diseases are favored by an acid condition, while others are not. Certain fungous diseases like "finger-and-toe" develop to a harmful extent only in acid soils, whereas other plant diseases like the potato scab are most serious in well-limed soils, and, hence, some acidity (pH 5.2) is desirable under many conditions for potato grow- ing. Aside from the cases known to have a direct relationship to plant diseases, there are a few special plants like the cranberry, blackberry, and watermelon for which a soil of at least slight acidity seems desira- ble and in some cases necessary for the best growth. It is not known just why these plants grow better on an acid soil, but it seems possible that in some cases plant diseases or malnutrition due to a lack of iron where the pH is too high may again be factors. The influence of soil reaction on the competitive powers of different species of plants to establish themselves and crowd out others needs to be recognized. For example, the common sorrel (Rumex acetosella) is sometimes said to grow best on acid soils. However, when grown under controlled conditions free of competition by other species of plants, it grows better on neutral than acid soils. It is commonly found grow- ing on strongly acid soils because there it usually meets with less com- petition from other weeds as well as crops. It would be found growing even better on the less acid soils were it not that here other plants grow so well and vigorously that they crowd out the sorrel. Undoubtedly, competition working in conjunction with soil acidity or alkalinity in the way just indicated has a marked influence on the powers of different species of plants for establishing and maintaining themselves under natural conditions, and thus has affected the charac- ter of the native vegetation found in many regions. Crop plants as well as plants in general vary greatly as regards the reaction of the nutrient medium in which they grow best. Some species of plants grow best in an acid medium, some in an alkaline medium, and others in a neutral or nearly neutral medium. The reasons for these reaction preferences are due to the direct and indirect influences of soil reaction just discussed. Emit Truog 43 SOILS ARE FRUGAL CUSTODIANS OF PLANT NUTRIENTS The question is frequently asked. Is there not great danger of loss of plant nutrients by leaching when fertilizers are added to soils? Or, What about fixation of the nutrient elements by the soil so that plants cannot use them? The answer is that good soils are frugal custodians of the plant nutrients naturally present in or added to soils. If that were not the case, soils would rapidly become so depleted of fertility elements by careless and exhaustive cropping, that a protective covering of crop plants, or even weeds, would not be produced. Under these conditions, erosion would be terrific and disastrous. The three availability categories of the nutrient elements Just how does a good soil perform the miracle of supplying plants with adequate amounts of nutrient elements and still prevent undue loss by leaching? It does this by forming combinations of the nutrient elements of varying degrees of solubility or availability. For conveni- ence of exposition and consideration, it may be said that nutrient ele- ments are held or stored in soils in three degrees or categories of avail- ability— namely, (a) readily available, (b) moderately available, and (c) slowly or difficultly available. To produce a good crop, it is necessary that the supply of readily available forms of nutrients be high enough so that much or most of what is needed by any crop can be obtained directly from these forms. However, it is a great mistake to assume that plants get no nutrients at all from the less available forms. Failure by some to recognize that plants obtain some share of their requirements from nutrient forms of all degrees of availability has caused much confusion in their thinking about this matter. Also, it should be recognized that there is a continual transformation of each nutrient element from forms of one degree of availability to forms of other degrees of availability, and that the main direction of this transformation is determined largely by the relative proportions of the various forms present. Figure 4 provides a schematic representation of the three availability categories of plant nutrients in soils. Transformation from one category to the others is represented by arrows in the connections between the 44 Mineral Nutrition of Plants categories, and the relative rate of possible transformation is indicated roughly by the width of the connections. Forms of potassium in relation to availability categories To serve as an example, the forms of potassium which make up the various categories of this element are given in the diagram. In this case, the exchangeable form makes up nearly all of the readily availa- THE THREE AVAILABILITY CATEGORIES OF PLANT NUTRIENTS IN SOILS J \ READILY AVAILABLE T MODERATELY AVAILABLE 1 SLOWLY AVAILABLE Figure 4. Schematic illustration of availability categories of plant nutrients in soils (exemplified with the element potassium) and trans- formation from one category to another. Arrows indicate the direction of transformations, and the wider the connection the more rapid the rate of transformation. ble category. Only small and rather insignificant amounts exist as water-soluble salts, such as nitrates and sulfates. Moderately available potassium (extracted with dilute acid) exists in the so-called "fixed state" as hydrous mica (illite) and biotite. It appears that at high levels of exchangeable potassium, some of this potassium gradually becomes transformed to a nonexchangeable or fixed form. Possibly in this transformation the montmorillonite which holds the exchangeable potassium goes over to hydrous mica or some Emil Truog 45 similar type of mineral. Since biotite contains ferrous iron which suffers oxidation on exposure to air and water, this mineral is quite readily subject to breakdown and release of its potassium. The potash feldspars and muscovite contain practically all of the slowly or difficultly available potassium. Muscovite, unlike biotite, con- tains no ferrous iron and, as a result, is tremendously resistant to weathering and release of its potassium. Potash feldspars, being dense crystalline silicates, weather slowly. This is particularly true of micro- cline. Although soils frequently contain 30,000 to 40,000 pounds per acre-plow-layer of muscovite and feldspar potassium, usually not more than five to ten pounds of this become available annually for crop use. However, if little or no plant growth is removed from land for several years and leaching is not too severe, there results a sufficient accumu- lation from this source to be a potent factor in supplying potassium to one or more subsequent crops. The so-called "resting of land" owes its virtues to processes of this kind. Let us now suppose a crop of corn is growing on the soil in question. What happens? There is a heavy drain on category A and, hence, its pressure for transformation to category B drops. As a consequence, the rate of transformation from B to A exceeds the reverse transformation and there is a gradual replenishment of A from B. After the growth of a heavy crop, it may take several months or more for the replenish- ment to run its course so that A and B are again in equilibrium. That is why the amount of readily available potassium found in a soil by test in the fall after the removal of a heavy crop is sometimes consider- ably lower than the amount found the next spring. It will be noted in the diagram that there is a narrow connection between categories A and C, and the arrows indicate some transforma- tion from C to A. However, the rate of this transformation is so slow that what may be transformed during a growing season is sufficient to supply only 5 to 10 per cent of crop needs. Thus it is apparent why the amount of potassium in A must be kept up to a certain level if a crop is to be adequately supplied. In practice it has been found that category A should contain about 200 pounds per acre-plow-layer of potassium to assure good yields of corn, alfalfa, and other general farm crops. Potatoes, sugar beets, and truck and garden crops require a level 46 Mineral Nutrition of Plants which is two to three times as high. Fortunately, the level of potassium in category A is easily determined by means of a chemical test. The amount of potassium in category B may be several times that in category A, but ordinarily during a growing season crops probably get no more than 10 to 20 per cent of their potassium needs from the former. The very important consideration in this connection is that during the time when crops are not growing, there occurs a flow of potassium from category B to A to make up some of what the previous crop has removed from A. Transformation of potassium from category B to C and the reverse of C to B is extremely slow. In fact, there is some question whether or not these transformations in the case of potassium take place at all, especially from B to C. Now, what happens when soluble fertilizer potassium is added to a soil in which categories A and B are at equilibrium? At first, most of this potassium rapidly goes into the exchange form of category A. This causes a disturbance of the previously established equilibrium between A and B, and as a result there occurs a flow of potassium from A to B until equilibrium is again established. That is why it takes more than the calculated amount of potash fertilizer to raise the level of potassium in A to some prescribed point. However, it is important to recognize that this potassium which is transformed to category B is not lost. It is simply stored more safely against leaching and reckless use by care- less cropping. Yes, the soil is a frugal custodian. What is the situation in virgin soils as regards levels of potassium in the three availability categories? Let us consider first the conditions under which the podzols are formed. Here, leaching is so severe that even a frugal soil is unable to stem the tide of relatively rapid loss of potassium from category A. This causes a continual flow from B to A, and, as a result, the level of potassium in categories A and B becomes low. The flow of potassium from C to A is so slow that virgin podzols in northern Wisconsin, containing less than 100 pounds per acre-plow- layer of category A potassium, have on the same basis 30,000 to 40,000 of this element in category C. This is further evidence that category C cannot be depended upon to furnish much of the current needs for profitable cropping other than forestry where the annual drain is low and only one crop may be removed in 50 years. Moreover, during this Emil Truog 47 time, the dropping of needles and leaves provides a continual return of nutrients to the soil. Even with this return, podzols in general are so low in category A and B potassium that soon after being brought un- der cultivation they require heavy applications of potash fertilizer. Going now to the slightly podzolized soils of southern Wisconsin, it is found that the supplies of potassium in categories both A and B are usually five to ten times as high as in podzols to the north, although the supplies in category C may be quite similar. This further em- phasizes the fact that transformation from category C to A is very slow. Because of the much greater supplies of category A and B potassium in the less podzolized soils, it is found that many of them may be cropped for 50 years or more without need of potash fertilization. The greatest supplies of category A and B potassium are found in the chernozems and other alkaline soils formed under limited rainfall. Here the build-up may be high enough to last for 100 years or more of cropping without the need of adding commercial potash fertilizer. (A more complete discussion of potassium in this connection may be found in references /, 2, and 7.) Thus far, nothing has been said about phosphorus, nitrogen, calcium, and other nutrient elements in relation to availability categories. In general, they follow much the same pattern as potassium, but, of course, the types of compounds involved may be quite different. Forms of calcium and magnesium in relation to availability categories Calcium and magnesium as regards category A are very similar to potassium, also being held there as exchangeable cations. In fact, unless the soil is quite acid or the pH is higher than 9, calcium and magnesium practically always make up the great bulk of the total exchangeable cation content of a soil; and, if the soil is quite acid, then calcium and magnesium are usually supplied in the form of lime to bring up the supply of these elements in category A to the desired level so as to attain a pH of at least 6.5. The calcium and magnesium of category A do not become fixed like potassium and revert to category B. In fact, relatively little category B calcium ordinarily exists, unless the exchangeable calcium still re- 48 Mineral Nutrition of Plants maining in distinctly acid soils be thus designated because of its rather low availability compared with much of that existing at pH 6.5. Mag- nesium like potassium tends to form secondary silicates to a much greater extent than calcium, and, hence, its greater tendency to go into forms which fall in category B. Because calcium feldspars and other calcium silicates and most of the primary magnesium silicates weather relatively rapidly, the supply of calcium and magnesium in category C is generally very much lower than potassium, especially in soils of the humid regions. That is one reason why in these regions it is generally necessary, from time to time, to add large amounts of calcium and considerable magnesium in the form of lime. Availability categories of phosphorus, nitrogen, and sulfur Since phosphorus exists in soils largely as the anion P04~ _, it cannot exist in category A in the same form as most of the calcium, magnesium, and potassium. It can, nevertheless, probably exist attached to aluminum and iron hydroxides and silicates as an exchangeable anion, and a small portion of this may possibly be considered as be- longing to category A. However, the main bulk of category A phos- phorus in soils well supplied in this respect exists as calcium phosphate. Experience teaches that for good yields in general farming, a soil should contain at least 50 pounds per acre-plow-layer of category A phosphorus. For truck and garden crops, double and treble this amount is required for best results. The forms of phosphorus which make up categories B and C have not as yet become well defined. Possibly, much of the phosphorus attached to aluminum hydroxide and the silicates can be considered as belonging to category B, and most of that attached to iron hydroxide and all existing as apatite as belonging to category C. Details regarding the manner in which phosphates and other nutrient anions are held and become available in soils are given in another paper of this sym- posium. Because nearly all of the nitrogen of soils is generally associated with organic matter rather than mineral matter, it might be expected that the pattern of availability categories for nitrogen would be radically Emil Truog 49 different from what it is for, say, potassium. This, however, is not the case. In order for the soil to be a frugal custodian, it appears necessary or desirable that the same general pattern be followed for all of the nutrient elements. The nitrogen of category A usually exists largely as fresh crop resi- dues and recently applied manure and nitrogen fertilizers. Very small amounts may be held in exchangeable form as NH4~. When organic matter accumulates for long periods as it has in chernozems, and mucks and peats, true humus is formed, and much of the nitrogen contained therein, possibly two-thirds or more, is of low availability and belongs to category C; the remainder falls in category B and largely provides the nitrogen for crops when these soils are brought under cultivation. Under continuous cropping, very little of the nitrogen added in the form of crop residues, manure, and ferti- lizer becomes transformed into category B or category C. However, when sod crops occupy the land continuously for several years, such transformation does take place to an appreciable extent. Thus, it is ap- parent that categories B and C with respect to supplies of nitrogen can only be built up in practical farming by keeping the soil in sod crops a good share of the time. Sulfur, as regards availability categories, follows a pattern quite similar to nitrogen when it is present as a constituent of organic matter. It also exists in mineral forms which fall into all three categories, but often to the greatest extent in category C. Availability categories of the minor nutrient elements The four minor nutrient elements — iron, manganese, copper, and zinc — which exist in soils primarily as cations follow much the same pattern as calcium and potassium. Contrary to this pattern, iron and manganese particularly when the pH is above 6.5, readily form the in- soluble higher oxides which belong to categories B and C. Boron fol- lows a pattern quite similar to nitrogen except that it changes rapidly to category C at moderate alkalinity, and again rapidly to A and B at high alkalinity. To be sure, many if not most of the details of the matter just dis- cussed are still to be worked out, especially with respect to the minor 50 Mineral Nutrition of Plants elements. However, enough is known to make it safe to repeat that a good soil is, indeed, a frugal custodian of all the nutrient elements. SOILS NEED ADEQUATE CAPITAL It is generally recognized that the success and survival of a business venture depends primarily upon adequate working capital and good management. In a sense, the same situation prevails in the case of a soil serving as a medium for plant growth. Here, the supply of readily available nutrient elements may be likened to working capital; the moderately available, to negotiable bonds and securities; and the slowly available, to fixed assets or investments in real estate. The farmer must, of course, provide the management. If a soil, located favorably as regards climate and surface drainage, is provided with an adequate supply of the essential plant nutrients in readily available forms, it is really astounding how satisfactorily crops will usually grow and how well unfavorable physical conditions of the soil will in time become corrected. In this connection the writer has in mind a specific case. In North Central Wisconsin there exists a large area of podzolic soil called Spencer silt loam. This soil is strongly acid throughout the profile, and is therefore low in its supply of available bases, particularly potassium. Also, the subsoil or B horizon is very compact and rather impervious to both water and roots. It was generally surmised that successful alfalfa culture on this soil would never be pos- sible, even if adequately limed and fertilized, because of the impervious and poorly drained subsoil. About twenty years ago a farmer by the name of James Asplin acquired some typical Spencer silt loam. Apparently sensing that the soil lacked a lot of something, Mr. Asplin proceeded to apply about twenty-five wagon loads (25 tons) per acre of wood ashes to his soil. The ashes were obtained from a nearby sawmill and probably contained about three per cent of potassium, 50 per cent of lime, and appreciable amounts of phosphorus and other mineral nutrients. Thus, approxi- mately 1500 pounds per acre of potassium in the form of the carbonate and an abundant supply of readily available lime and other mineral nutrients were provided. In other words, the soil was provided with adequate working capital. Emil Truog 51 Following the application of the ashes, crops began to grow with great vigor. Even alfalfa became a successful crop on a soil whose highly acid, impervious, and poorly drained subsoil would seem to rule out this crop for all time. Just lately, an alfalfa plant, which among many others has persisted for about fifteen years in a subse- quently unplowed corner of one of the fields treated with ashes, was carefully dug up so its root development might be examined. The root system of this plant was found to be both deep and ex- tensive. Large tap-roots had gone to a depth of over four feet. Gnarled knots found on the roots indicated clearly that at certain stages the growing points of some of the roots probed about persistently in search for a tiny opening which would allow further downward penetration. Undoubtedly, the vigor in the plant needed for this relentless probing and development was engendered by the liberal supply of plant nu- trients. Extensive field tests in recent years show conclusively that with adequate liming and fertilizing, Spencer silt loam becomes a very good medium for the growing of alfalfa and many other crops. Under this treatment and a rotation which includes a tap-rooted legume like alfalfa, there is little question that in time this soil can be vastly im- proved. When these large tap roots die and decompose, the subsoil will be provided not only with organic matter, but also, and this is probably even more important, with large pores for aeration and drainage. Of course, there are some soils which, because of shallowness, high water table, or other physical defects, cannot be made productive by applying lime and fertilizer. Then there are the saline soils — soils which because of restricted leaching contain such an abundance of salts (some of these may be actual nutrients) that a toxic condition for plant growth is created. To carry out the simile used previously, it may be said that such soils are over capitalized and, for amelioration of this condition, require special treatment. Wherever soils have suffered severe leaching, they usually lack an adequate supply of plant nutrients for good crop production. However, when this lack is corrected, the favorable results are often phenomenal. The present high productivity of podzolic soils in the Scandinavian countries furnishes proof of this. 52 Mineral Nutrition of Plants DETERMINING A SOIL's FERTILIZER NEEDS AND FITNESS FOR PLANT GROWTH If a soil has been properly classified much will be known with re- spect to its management needs and crop adaptability. If it is a podzol, then we know it will be acid and low in available supplies of most of the plant nutrients. Soon after being brought under cultivation, it will need lime and fertilizer, particularly potash, if legumes like clover and alfalfa are to be grown. If it is a fine-textured podzol, then we also know that the subsoil will probably be rather tight and impervious. This may require special treatment. On the other hand, if the soil is a chernozem, that will immediately stamp it as being well supplied with nutrients and an excellent medium for plant growth. It will need no lime and probably no fertilizer for many years. The first fertilizer needed will probably be phosphate. It should be an excellent wheat and alfalfa soil. When fertilizer treatment of soils becomes necessary because of either a low level of available nutrients to start with or because of exhaustive cropping, then recognition that plant nutrients in soils exist in several categories of availability, and that transformation from one category to another takes place at a rate dependent on a number of specific conditions, will be basic to any satisfactory program of de- termining and interpreting the nutrient status of the soils. Originally the nutrient status was determined for the most part by total analysis. Then when the results did not correlate at all with crop growth and fertilizer response, attempts were made to determine the more readily available portions of the nutrient elements. Because existing knowledge of the actual forms in which the nutrients exist in soils and of the man- ner in which they change from one form to another was meager indeed, progress was slow, and faith in the eventual success of soil testing was abandoned by many workers in this field over a long period of time. Advances in knowledge during the past 25 years with respect to exchangeable bases, to types of minerals and compounds that exist in soils, and to conditions (including in particular the pH of the soil) which determine rate and type of transformation from one form to another have all contributed greatly to the development of better and Emil Truog 53 better soil tests and more satisfactory interpretation of the results thus obtained. Today, soil testing is carried forth on a vast scale in many states and countries. In Wisconsin during the present year, several hundred thousand soil samples will be tested for pH, readily available phosphorus and potassium, and in some cases for available boron and other nutrient elements. The determination of the pH of soils is now both simple and ac- curate. The result tells immediately whether or not it is advisable to add lime for the most satisfactory growth of various crops. It also tells when a soil is so highly alkaline that special treatment is needed to counteract the condition. Recognition that three categories of availability of each plant nutrient should be considered and that transformation from one category to another takes place is of tremendous help in the interpretation of re- sults of soil tests for potassium. For example, in the case of this ele- ment, it is now quite generally recognized that for the satisfactory production of general farm crops, a minimum level of 200 pounds per acre-plow-layer of readily available (category A) potassium should be present. However, if this drops, say to 100 pounds through cropping, it is a much more serious matter in a strongly podzolized soil of north- ern Wisconsin than in a mildly podzolized soil of similar texture in southern Wisconsin, because in the former there will be much less category B potassium present which will change to the A category with crop feeding. Also, if it is desired to raise the level of category A potassium to 200 pounds in both cases, it will take considerably more potash fertilizer in the strongly podzolized soil because of more rapid and greater transformation to the naturally depleted B category. Although interpretation of results of soil tests for the other nutrient elements may vary considerably from that for potassium, nevertheless, as it now appears, the general pattern should be quite similar. To be sure, much more research is needed in connection with methods of determining the amounts and forms of the various nutrient elements present in soils in the three availability categories. With the knowledge and special analytical instruments now at hand and being developed, further progress in this field is certain to be rapid. Thus, soil fertiliza- tion and management to the end that soils be made still better mediums 54 Mineral Nutrition of Plants for crop growth will be carried forth with ever-increasing confidence and success. SUMMARY Soils may be considered and studied from two standpoints. First, they may be looked upon as being natural entities or bodies, just like rock formations or rivers, or even plants and animals, and then studied with respect to their origin, morphology, and classification without any specific regard to their agricultural use. This kind of consideration of soils constitutes the science of pedology, which provides a sound and scientific basis for soil classification and mapping. Although pedology is not concerned directly with soil fertilization and management, it does point the way to the basic system of soil classification needed in determining proper land use and crop adaptation. In the second place, soils may be looked upon solely from the stand- point of being media for plant growth. As the title of this paper sug- gests, most of the discussion has been given from this standpoint. It was pointed out that soils are not only three-phase systems in the ordinary sense of the physical chemist, but have in addition a "living phase" represented by bacteria and other microorganisms whose num- ber in a thimbleful of fertile soil may exceed the number of people on this earth. The soil is, indeed, an intricate dynamic system, and to understand how a fertile soil may function as a well nigh perfect medium for plant growth requires much study and considerable imagi- nation. Not only must a good soil provide anchorage and aeration for plant roots, store and deliver enormous amounts of water to growing plants, but it must also serve as a frugal custodian of a dozen or more nutrient elements. If a soil were not frugal with its resource of nutrient ele- ments, it would soon become so depleted of these elements that destruc- tion by erosion would be certain and rapid because a protective vegeta- tive cover would then fail to grow. The manner in which a soil func- tions as a frugal and judicious custodian of nutrient elements and how the amounts of these in various forms should be considered and may be determined is explained in considerable detail. The unfailing de- pendability of a soil serving as a medium for plant growth is beauti- fully summed up in verse by David Grayson as follows: Emil Truog 55 Why risk with men your hard-won gold? Buy grain and sow; your Brother Dust Will pay you back a hundredfold — The earth commits no breach of trust. REFERENCES 1. Attoe, O. J., and Truog, E., Soil Sci. Soc Am. Proc, 10:81-86 (1945). 2. Attoe, O. J., Soil Sci. Soc. Am. Proc, 11:145-149 (1946). 3. Baldwin, Mark, Kellogg, Charles E., and Thorp, James, Soils and Men, U. S. Dept. Agr. Yearbook (Washington, 1938) 979-1001. 4. Byers, H. G., Kellogg, Charles E., Anderson, M. S., and Thorp, James, U. S. Dept. Agr. Yearbook (Washington, 1938) 948-978. 5. Bryan, O. C, Soil Sci., 15:23-35 (1923). 6. , Soil Sci., 15:375-381 (1923). 7. Evans, C. E., and Attoe, O. J., Soil Sci., 66:323-334 (1948). 8. Jenny, Hans, Factors of Soil Formation (1st ed., New York, McGraw- Hill Book Co., 1941). 9. Kellogg, Charles E., The Soils That Support Us (New York, The MacMillan Co., 1941). 10. Magistad, O. C, Soil Sci., 20:181-225 (1925). 11. Truog, E., Soil Sci. Soc. Am. Proc, 11:305-308 (1946). 12. , Soil Sci., 65:1-7 (1948). CHAPTER O The Activities of Cations Held by Soil Colloids and the Chemical Envi- ronment of Plant Roots C. EDMUND MARSHALL T he concept of the soil solution encounters difficulties as soon as ionizing colloidal constituents contribute appreciably. In the author's opinion, these difficulties become insurmountable if the soil solution is defined either according to Burd's recent dictum (2) in which the contributions of the colloidal constituents or ionizing surfaces are excluded, or according to earlier views identifying the removable soil solution (obtained by displacement or by pressure) with the soil solu- tion in situ. On the other hand, a frank recognition of the complexities involved, together with the utilization of modern methods for de- termining cationic activities in colloidal systems, opens up the possi- bility of their resolution. THEORY OF THE SOIL SOLUTION In what follows, the soil solution will be defined as the complete external aqueous chemical environment of the plant root. This is a macro concept, that is, it involves the mean chemical environment for each ionic and molecular species present. In actual soils variations un- doubtedly occur from place to place on a single root hair, but we are dealing here with averages taken over large numbers of points of con- tact and immense multitudes of ions. It is also a static concept, avoid- ing reference to changes with time except in so far as these may be 58 Mineral Nutrition of Plants determined experimentally by activity measurements. This is not so serious a weakness as might be supposed, since any future kinetic treatments of ionic transfer between soil and plant will inevitably need both a firm point of departure and a practical experimental method for observing macro changes with time. Both are now available. Some simplifying assumptions will be made regarding the nature of the soil systems. Sparingly soluble salts such as calcium carbonate, tricalcium phosphate, and calcium sulfate will be omitted from con- sideration. The arguments used below can readily be modified to include them. The soil system will be assumed to consist of (a) an inert skeleton, (b) negatively charged soil colloids holding cations in a diffuse ionic atmosphere, (c ) soluble salts, and (d) water. Consider now a large mass of such an idealized soil at a given mois- ture content. Suppose that a small amount of solution is forced out by pressure, the removal causing no appreciable change in the original soil. Then the relationship between the concentration of soluble elec- trolytes in the solution thus expressed, and in the soil mass, is governed by the Donnan equilibrium. It can therefore readily be determined. This means that we can define the conditions under which the con- tribution of the ionizing colloids becomes important and can evaluate the relationship between the soil solution as defined in this paper, the "Burdian" soil solution, and the solution expressed from the soil. To do this we must assume that single cationic activities have mean- ing and are measurable and must use certain simplifications in order to get an approximate over-all picture. Consider potassium-saturated soil in the presence of potassium chlo- ride solution and let a,- represent the activity of the potassium ions associated with the soil colloid, aVt the activity of potassium ions in the "Burdian" soil solution (i.e., due to potassium chloride alone), and aE the activity of potassium ions in the expressed fluid which is free from soil colloids. According to the Donnan equilibrium we then have (ac + aB)aB =z (aE)2 (1) assuming that chloride ions have the same activity coefficients as potas- sium ions. What we need is the relationship of aVj to an when ac takes on different values in relation to av,. It can easily be seen that for finite C. Edmund Marshall 59 positive values of ac, aE is always greater than aB. Let aE = qaB. Then by substitution it is easy to show that aJaB = q2 — 1 (2) giving a well-defined value of q or aE/aB for each value of ac/aB. The curve is a parabola, and we can see the situation at a glance (Figure 1). Under the above assumptions the curve will apply to any electrolyte °z/°i lb 12 10 ^ 0 1 .^ J 0°— J A 8 ^ -G\°S / •» ja - B 6 / y r,o< ;>?. 4 CO <- CO*. ncoj^z Coc 2 0 2 0 4 0 6 0 8 0 IC )0 1 BO U iO 1 50 °C/°B Figure i. The Donnan equilibrium between soil colloids and expressed solution. For symbols see Table I. whose anion and cation have the same valency, and if several such are present simultaneously each can be dealt with. Thus, if we know ac, the cationic activity of the soil colloid, we can calculate how large the salt content will be for any chosen ratio aE/aB. Alternatively, for any value of aB, corresponding to the "Burdian" soil solution, we can de- termine how large the cation activity of the colloid needs to be for q to assume any value greater than 1. Some interesting limiting cases arise. Obviously, if ac is very much less than aB, the "Burdian" soil solution becomes practically identical with the expressed solution and both are indistinguishable from the 60 Mineral Nutrition of Plants whole soil solution as here defined. Where appreciable colloid is present, it takes a high salt concentration to achieve this. Where the colloid content is very low, a much lower salt concentration will suffice. This explains why those who have worked with plant growth problems in highly saline soils and those who have had similar experience only in extremely light sands have been satisfied with a concept involving simple solutions of ordinary electrolytes. However, the great preponder- ance of productive agricultural soils are neither highly saline nor de- void of ionizing soil colloids. Turning to the case in which a(. is much greater than aB, we arrive at the formula for membrane hydrolysis, in which the compound under consideration is, in the above example, potassium hydroxide. Experi- mentally this is very simple, since (ac -f- aB) is given by the cationic activity of the whole soil system, aB by the hydroxyl ion activity of the whole soil system, and aE corresponds to the potassium hydroxide in the expressed liquid. Thus, if we know the pH of the soil system and its cation activity, we can immediately determine the composition of the expressed liquid. The strength of the soil colloid as an electrolyte comes into this picture only through ae, the cation activity of the colloid under the given moisture conditions. The expressed liquid will be alka- line, hence this equilibrium will be very sensitive to the presence of weak acids, such as carbonic, in the system. It is easy to see that such an acid will greatly increase the potassium activity in the expressed liquid. It does not do this, however, primarily by competition with the soil col- loids, which are much stronger acids than carbonic, but indirectly through its effect on the Donnan membrane hydrolysis. Similar considerations may be applied to other cationic constituents of soil colloids; as for instance, calcium-saturated soils in equilibrium with calcium chloride. For this case the equation corresponding to (2) is ac/aB = qz - 1, graphically represented also in Figure 1. Here again we reach the limiting condition for membrane hydrolysis when ac is very large in comparison with «B. The hydrolysis can be calculated from the equilibrium with respect to calcium hydroxide, using the appropri- ate Donnan equation. It will be a sensitive function of the pH of the soil system, since the hydroxyl ion concentration comes in as the second power. C. Edmund Marshall 61 Table I illustrates the interconnections of the three concepts of the soil solution. The three ranges chosen in each case for ac cover the cases most likely to be encountered. Three chloride concentrations are in- cluded under aY> and in addition the situation in the absence of salt is examined, as it would be at the neutral point. Here the hydroxyl ion activity is io-7 and a,. -\~ av, is practically identical with ac. aE then rep- resents hydroxide expressed from the soil. The effect of pH on the TABLE I The Operation of the Donnan Equilibrium in Soil Systems Cation* ac aB ac + aB aE q — aE/aB Potassium io~2 io-' o.oioo 3.16 X io~° 0.0101 1.005 X I0~ 10.05 0.0 1 10 3.32 X 1 o~ 2.32 0.0200 x .41 X io~ 1 .41 0.00 1 00 1 .00 X io" 0. 001 01 1.005 X I0~ 10.05 0.00110 3.32 X 10 3-32 0.00200 1. 4 1 X 10 1. 4 1 Calcium io~J io-' 0.00100 2.15X io~6 0.00101 4.66 X io- 4.66 0.00110 2.24 X io~ 2.24 0.00200 1.26 X 10 1.26 0.000 1 00 1 .00 X 10 0.000101 4.66 X 1 o- 4.66 0.000110 2.24 X 10 2.24 0.000200 1.26 X 1 o~ 1.26 *<x -4 10 0.016 loam 3=2 M 3X -3 10 0 161 16 4X -4 10 0.015 (beidellitic) 3:I 58 oX -3 10 0 136 35 8X -4 10 0.017 Cecil clay 3:8 6.8X -3 10 0 267 I=i 4X -4 10 0. 131 (kaolinitic) 3:4 12 1 X -3 10 0 258 22 1 X -4 10 0.094 3:2 19 9X io-3 0 211 32 8X —4 10 0.070 measured activity to the total concentration of the ion concerned. In true solutions it is known as the activity coefficient. The following con- clusions then emerge from an examination of Tables I and II together. 1. In spite of the fact that the clay colloids are weakly ionized; the cationic activities of salt-free soils in the moisture range where plants actually grow are significant. Even in the case of calcium we have activ- ity values of the order of 3 X io~3, which is considerably greater than would be given by calcium carbonate in equilibrium with ten times the carbon dioxide pressure of ordinary air. 2. The actual activities obtained depend on the nature of the soil col- loid present as well as on the amount. 3. The cationic activities are a sensitive function of the soil-water C.Edmund Marshall 63 ratio. This means that the chemical environment of the plant root changes greatly with changing moisture conditions in the soil. 4. Since different cations are ionized to different degrees, it becomes exceedingly important to determine all the factors involved. In a nat- ural soil system with a mixture of exchangeable cations present, the relative activities may be quite different from the relative exchange quantities. 5. The expressed soil solution will be considerably different from the soil solution in situ at the low salt contents of nonsaline soils containing appreciable amounts of soil colloids. In many cases the cationic environ- ment of plant roots will be provided chiefly by the soil colloids. In view of these conclusions the detailed study of the electrochemical properties of soil colloids becomes of great importance to plant nutri- tion. The main results thus far obtained are reviewed below. THE CATIONIC ACTIVITIES OF SOIL COLLOIDS The colloids of soils may be divided into two main groups, inorganic and organic. In the inorganic group, the clay minerals are of predomi- nant importance in contributing cations. The organic colloids are domi- nated by humified materials of an electronegative character having, in general, higher capacities for cation exchange than the clays. In a ma- jority of fertile soils, the content of organic matter is less than that of the clay minerals. The electrochemistry of these two groups of soil colloids had, until the advent of clay membrane electrodes, been developed through pH measurements and conductivity determinations. Potentiometric and conductometric titrations of the free colloidal acids, generally purified by electrodialysis, had led to the conclusion that both the clays and the humic materials were genuine colloidal acids, owing their acidic char- acter to inherent structural features and not acquired by adsorption of soluble acids from solution. Many attempts were made to compare them with weak soluble acids by ascribing to them approximate dissociation constants. Evidence on the exchange properties of these soil colloids also con- tributed to our knowledge. Exchange isotherms were obtained by meas- uring analytically the exchange equilibria for different proportions of 64 Mineral Nutrition of Plants colloid and electrolyte. These data, like those obtained potentiometri- cally, were generally interpreted on the assumption that all exchange cations of a given kind on a given colloid were held by the same bond- ing energy. Some evidence at variance with this simple concept grad- ually accumulated but was not regarded as sufficiently general to in- validate it. Three main items may be mentioned, (a) Certain potentiometric titration curves of clays showed more than one inflection, {b) The last traces of exchangeable hydrogen were found, in general, to be extremely difficult to replace by metallic cations, (c) Hysteresis effects were found in certain exchange reactions; that is, the equilibrium was different when approached from different sides. The logical consequences of a variation in the bonding energy of a given cation upon a given soil colloid were first adequately discussed by Jarusov (4), who presented clear experimental evidence that the replaceability of metallic cations varied with the degree of saturation. He emphasized also the importance of the nature of the accompanying ions in determining the replaceability of a given cation, thus fore- shadowing the important work of Jenny and Ayers (5) on the comple- mentary ion principle. The work summarized below was made possible by the development in the Missouri Experiment Station, of membrane electrodes which, acting in a somewhat similar fashion to the glass electrode, render the determination of the activities of single cations feasible. The sequence of events was briefly as follows. Thin plates cut from crystals of cation exchange minerals were first employed. Below a certain concentration they functioned as electrodes sensitive to a variety of cations. Their preparation was difficult. The next materials tried were clay films produced by the evaporation of colloidal suspensions of high base-exchange clays. By preliminary heat treatments these membranes were found to acquire good mechanical stability and at the same time their electrochemical properties were im- proved (//). In the case of the hydrogen montmorillonite, it was found that membranes could be prepared which were either sensitive both to mono- and divalent cations (pretreatment below 450 ° C.) or which were sensitive only to monovalent cations (pretreatment above 4500 C. Edmund Marshall 65 C). These membranes were used in the determination of the activities of K+ (12), NH4+ {is), and Na+ (75) when these cations were added as hydroxides to carefully purified acid clays. Thus, complete titration curves, involving both the hydrogen ion and the added cation, were obtained for the four clay minerals: montmorillonite, beidellite, illite, and kaolinite. Divalent cations were then considered, calcium being the first inves- tigated. Accurate procedures were established for the determination of calcium and magnesium activities (10, 14)', these were later extended to barium (j). Results comparable to those given by the monovalent cations have now been obtained for montmorillonite, beidellite, illite, and kaolinite using magnesium, calcium, and barium (_?). In this way very interesting evidence on the ionization of the clays was afforded and, as mentioned previously, its extension from homi- onic clay suspensions to homionic soils offers no difficulties. Attention was then directed to a more difficult problem, namely, the simultaneous determination of monovalent and divalent cations present together upon the clay. Calcium and potassium were investigated first. By means of high temperature selective membranes, the potassium activities in such mixtures were readily and accurately determined. The calcium estimation was subject to a considerably greater error but suffi- ciently reliable information was afforded to give an over-all picture. These potassium-calcium relationships have now been established for the following clay minerals: two montmorillonites, a beidellite, two illites, a kaolinite, a halloysite, and attapulgite (1,7). In a similar way potassium-magnesium relationships are under investigation. Work is also under way on similar evidence for the different fractions of soil organic matter. In reviewing the information already available, it has been apparent for some time that the complex nature of the complete titration curves of the clay minerals indicates that they hold single cations with a con- siderable range of bonding energies. It has proved possible to utilize the activity measurements to calculate these bonding energies. Thus, in addition to the pH curve and the cationic activity curve, we may derive also a bonding energy curve for each cation upon each clay. Where two cations are concerned we may express their mutual effects 66 Mineral Nutrition of Plants in terms of shifts in the bonding energy curves, comparing for instance, potassium-calcium with potassium-hydrogen systems and calcium-po- tassium with calcium-hydrogen systems. Evidence obtained with single monovalent cations The complexity of clay titration curves may be illustrated by the case of 2.8 per cent hydrogen bentonite and sodium hydroxide (Figure ph 900 800 700 AF 600 (calories 500 400 300 200 100 0 per equivalent ) 1 A 1 By c 1 / / t / J « 10 20 40 60 80 100 120 Miiliequivolents Na OH per 100 q cloy 140 Figure 2. Clay titration curves for 2.8 per cent Wyoming bentonite with sodium hydroxide. A: pH titration curve. B: Sodium ion activity plotted against base added. C: Mean free energy of sodium ions plotted against base added. 2). Three curves are presented. A is the ordinary pH titration curve, characteristically smooth and with a single reasonably well-defined in- flexion point. Curve B relates the sodium ion activities to the amount of base employed. Its complex characteristics are well established for C. Edmund Marshall 6j this combination of clay and cation. C is a cationic free energy curve. At each point the free energy value is calculated from the actual activity and the total concentration of the cation concerned. The formula may be written (AF) cat ion = RT ln Nation/* cation • The reciprocal of the cationic fraction active is therefore used in calcu- lating (AF)Na, since aNa (measured) /VNa (total) is the fraction active for sodium. Corresponding to the sharp changes in slope in curve B, we find regions of rapidly changing free energy in curve C. The range, in mean free bonding energy of the sodium, runs from about 500 to 800 calories per mole ion. Similar curves are found for potassium and ammonium although the extremes are not quite so far apart. As regards the other three clays: beidellite shows many of the features of montmorillonite but with a considerably smaller fraction active corresponding to greater free energy values; illite gives somewhat simpler activity and free energy curves and high free energy values; kaolinite resembles illite in its curves, but the actual free energies are the least of any of the four clays. In making comparisons of one clay mineral with another, it is important to keep the total cation concentration approximately con- stant, since for each clay the fraction active and, hence, the cationic free energy varies considerably with clay concentration. A series of values of the fraction active is given in Table III for definite stages in the neutralization of the acid clays with bases. The table does not, of course, display all the variations in the fraction active shown by the complete curves. The following conclusions may be drawn. t. If the state of each cation at the inflexion of the pH curve is taken as standard, then with decreasing degrees of saturation the fraction active falls, the rate of decrease being rapid down to 80 per cent satura- tion and slower between 80 per cent and 30 per cent. The bonding energy in the latter range runs from 20-50 per cent higher than that at the inflexion. 2. The fact that the fraction active is considerably less than unity sug- gests at first sight that we are dealing with weak electrolytes. However, the ionization is not appreciably changed by adding soluble salts with 68 Mineral Nutrition of Plants a common cation (12, /]). The analogy with weak acids and bases therefore breaks down. These colloidal electrolytes do not readily adapt themselves to classifications based ©n the behavior of small molecules in true solution. 3. Kaolinite stands somewhat apart from the other three clays in its ionizing properties. The acid clay is much less dissociated than hydro- TABLE III Cationic Fractions Active for Four Clays at Definite Stages of Neutralization* Percent- Fraction A :tive at the Follow- age of Equiva- ing Stages of Neutralization Concen- lence, Clay tration m.e./ioo Cation 50% 75% 100% Montmoril- 2.8 100 sodium o-377 0.258 0.381 lonite 3-3 100 potassium 0.295 0.271 0.297 3° 100 ammonium 0.264 0.249 0.245 Beidellite 4.0 70 sodium 0.096 0.092 0.131 (Putnam) 4.6 70 potassium 0. 122 0. 105 0.149 5.0 70 ammonium 0.053 0.067 0.098 Illite 10. 0 28 sodium 0.073 0.076 0.123 10. 0 28 potassium 0.144 0. 121 0.155 10. 0 28 ammonium 0.144 0.130 0.134 Kaolinite 10. 0 3 sodium 0.248 0.264 0-336 10. 0 3 potassium 0.195 0.238 0.326 10. 0 3 ammonium 0.234 0.255 0.252 *Data from references 12, 13, 15 gen montmorillonite, hydrogen beidellite, or hydrogen illite, yet the salts with monovalent cations are more dissociated. A structural reason can be suggested for this difference. Kaolinite has an electrostatically neutral framework, with ionization probably restricted to broken edges of the sheets. The other clays have negatively charged framework struc- tures, the charges being distributed over the cleavage surfaces of the sheets. These charges are neutralized by mobile cations whose activities we measure. C. Edmund Marshall 69 Evidence obtained with single divalent cations In comparing results obtained for divalent cations with those for monovalent, certain requirements as regards energy must be kept in mind. Suppose that two identical, neighboring, exchange spots release the same amount of energy when combining with two monovalent cations as they do with one divalent cation. The bonding energy per equivalent is therefore the same in the two cases. However, in terms of ions, the bonding energy per divalent cation is twice that per mono- valent cation. From the logarithmic formula relating bonding energy per mole ion to the fraction active, one can readily see that the fraction active of the divalent cation would be the square of the fraction active of the monovalent cation. It is therefore to be expected that divalent cations will show very low ionization from silicate surfaces. Thus, if the fraction active for a certain monovalent cation is 0.1, then that for a divalent cation releasing the same energy per equivalent should be 0.01. This is often found experimentally to be the order of magnitude of the relationship, allowing, of course, for variations due to cationic in- dividuality as it may show itself in hydration or in geometrical relation to the silicate surface. Figure 3 gives the curves, analogous to those of Figure 2, for hydrogen montmorillonite titrated with calcium hydroxide. The activity of cal- cium is seen to be almost constant over a considerable range, extending from about 30-70 per cent saturation with base. Then it rises very rapidly through the point of equivalence. This general behavior is found for all four clays (montmorillonite, beidellite, illite, and kaolinite) and for the three cations (magnesium, calcium, and barium) with some variation in the extent of the flat position and the subsequent rise. The actual values of the fraction active are summarized in Table IV. Since calcium is the major exchange cation in most fertile soils, a great deal of practical significance emerges from a study of these figures. Acid soils, if less than 70 per cent saturated with calcium, may be expected to give a very low calcium activity. The lime required to bring the saturation up to 70 per cent will make relatively little differ- ence as regards the chemical environment of the plant root. Between 70 and 100 per cent saturation the addition of lime greatly increases the jo Mineral Nutrition of Plants calcium activity. This picture, of course, may need modification for soils in which organic matter plays a predominant role, but it accords very well with practical experience on soils high in clay. The compari- son of kaolinitic soils with those dominated by members of the mont- morillonite group is also extremely interesting. 1300 1200 AF 1100 (calories, „ per 1000 equivalent ) 900 800 700 600 500 A / / *' B *- ■• — — • -♦— C 1 I 20 40 60 80 100 120 Milliequivalents Co(OH)2 per 100 q.clay 140 0*10 Figure 3. Clay titration curves for 1.07 per cent Wyoming bentonite with calcium hydroxide. A: pH titration curve. B: Calcium ion activity plotted against base added. C: Mean free energy of calcium ions (per equivalent) plotted against base added. Under any given set of circumstances the actual value of the fraction active determines the mean bonding energy of the cations present and, hence, controls the ease with which they can be exchanged for other cations. It is entirely possible to have two soils, dominated by different clay minerals, which would give the same calcium activity but different C. Edmund Marshall 7i bonding energies. For very small exchanges against hydrogen or other cations these would behave alike, but for moderate or large exchanges considerable differences would be apparent. Thus, in order to interpret the behavior of soils with regard to plant growth in the field, we should employ both the activities and the fraction active at various moisture contents. In addition we need control experiments designed to show TABLE IV Cationic Fractions Active or Four Clays at Definite Stages of Neutralization Using Divalent Cations Percent- Fraction Active at the : Follow- age of Equiva- ing Stages of Neutralization Concen- lence, Clay tration m.e./ioo g. Cation 50% 75% 100% Montmoril- 1 .04 100 magnesium 0.0122 0.0086 0.0085 lonite 1 .07 100 calcium 0.0175 0.0172 0.0530 1.07 100 barium 0.0036 0 . 0063 0.0235 Beidellite 4.8 70 magnesium 0.0059 0.0035 0.0068 (Putnam) 5.0 70 calcium 0.0040 0.0035 0.0081 3-4 70 barium 0.0036 0.0045 0.0109 Illite 4.9 28 magnesium 0.049 0.030 0.023 4.9 28 calcium 0.048 0.032 0.034 4-9 28 barium 0 . 0036 0.0027 0.0087 Kaolinite 9.0 3 magnesium 0.0097 0.0057 0.062 9.0 3 calcium 0.0134 0.035 0.092 9.0 3 barium 0.0174 0.065 0. 106 Note: The values for the equivalence have been arbitrarily chosen to correspond with those of Table III. Actually the inflexions on the pH curves are somewhat different for divalent cations than for monovalent. Data from reference j. how far the living root changes its external ionic environment as growth proceeds. Evidence obtained with one monovalent and one divalent cation Since we have perfected membranes of two types, those sensitive only to monovalent cations and those sensitive to all cations, it has proved possible, within limits, to determine one monovalent and one 72 Mineral Nutrition of Plants divalent cation in a mixture (7). This was regarded as particularly important because of the vast literature dealing with potassium-calcium relationship in plant nutrition. To clarify the potassium-calcium situa- tion in soil colloids naturally seemed the first step. If each of the cations concerned were held with a single bonding energy, then in a clay colloid saturated with a mixture of two cations we should anticipate that the fraction active for each should be constant. The activity for each cation would be directly proportional to the total amount present. On the other hand, with a range of bonding energies for each of the cations concerned, a variety of situations may arise, depending upon their quantitative relationship. Thus the fraction active of the one cation will become dependent upon the amount of the second ion present, as well as on its nature. In this sense one some- times refers to the effect of one ion upon the activity or the fraction active of another. Actually, no direct effect of the nature of a steric hindrance, etc., need be present. The mutual relationships found are consequences of the energy differences of reactive spots on the silicate surfaces. The importance of determining one cation in presence of another was realized as soon as membranes sensitive only to monovalent cations had been perfected. Measurements of potassium ion activities were made in a series of Putnam clay systems with differing proportions of calcium and potassium (12). It was concluded that in the middle range of base saturation (pH 5.5-57), there was little effect of calcium on potas- sium. We now know that this result is a special case and that it was largely due to the particular proportions employed. In general, as we shall see below, the potassium ion activity and the fraction active are greatly influenced by the proportion of calcium. The total information on these relationships is now considerable, the following list of clay minerals having been investigated; two ben- tonites, Putnam clay (beidellite), two illites, halloysite, kaolinite, and attapulgite. One of the bentonites was reported on by McLean and Marshall (7), the other systems have been investigated here by S. A. Barber (Ph.D Thesis, 1949). The complete results, including studies of magnesium-potassium relationships by E. O. McLean, will be pub- lished elsewhere. C. Edmund Marshall 73 Three sets of measurements were carried out on each clay system. By titrating the acid clays with potassium hydroxide and with calcium hydroxide and determining potassium and calcium activities, two bases of comparison were established. Then mixed potassium-calcium sys- 2 2 2 0 i e **- I ^ ■J y 1 6 S / r 1 4 a. « io3 1 2 K-Co 1 0 8 6 4 2 K- H 0 2000 £p 1800 K- H t calories per 1600 equivalent ) f 1400 1200 . < K- i < Ca — -- v. N. 1000 *1 20 40 60 Percentage K 80 100 Figure 4. Comparison of potassium ion activities and mean free energy values in potassium-hydrogen and potas- sium-calcium clay systems (2.5 per cent Putnam clay). terns were prepared (generally at a total saturation with bases of 100 per cent as judged by the inflexion on the pH curve), and the calcium and potassium ion activities were measured. Thus, complete informa- tion on the energy relationships of (a) hydrogen in presence of potas- sium 01 calcium, (b) potassium in presence of hydrogen or calcium, 74 Mineral Nutrition of Plants and (c) calcium in presence of hydrogen or potassium can be worked out. The data obtained under (c) are limited by the difficulty of measur- ing very low activities of divalent ions in the presence of much higher activities of monovalent. 1 2 1 | A i 1.0 / '/ / / 3 Ca-Kj '/ 7 // 1 OCQ * I04 .6 5 // 1 I lea- H 3 2 > 1/ » "*" --. 1800 Co- > 1 -K Ca-H < 1 i — AF (Colories Vv per l,uu equivalent ) -~*~ 1 1200 1000 Percentoge Co Figure 5. Comparison of calcium ion activities and mean free energy values (per equivalent) in calcium-hydrogen and calcium-potassium clay systems (2.5 per cent Putnam clay). Figures 4 and 5 show how the mean bonding energy of potassium is affected by the presence of calcium replacing hydrogen and how that of calcium is affected by the presence of potassium replacing hydrogen. The particular clay used was Putnam (beidellite). It is evident that calcium greatly lowers the energy with which potassium is held. This C. Edmund Marshall 75 result was obtained in varying degrees with all the clay minerals ex- amined. The influence of potassium on the bonding energy of calcium depends greatly on the relative amounts of these cations and on the clay mineral under investigation. The montmorillonite reported on by McLean and Marshall afforded Ratio 12 Ak//,ca O/lOO 10/90 20/80 30/70 % Exchangeable K / % Exchangeable Ca Figure 6. The potassium-calcium activity ratios for vari- ous proportions of the two exchangeable cations on a variety of clay minerals. a clear-cut case of the depression of calcium activity with increasing potassium and, hence, of an increase in the calcium ion mean bonding energy. The other clays are less regular and may show an increase over one part of the range and a decrease over another. A general review of these potassium-calcium relationships may be obtained by plotting the ratio of the activities against the ratio of the j6 Mineral Nutrition of Plants total exchangeable quantities. This has been done in Figure 6 in which the activity ratio K/Ca is plotted vertically and the concentration ratio horizontally. It covers the range from 70-100 per cent calcium and 30-0 per cent potassium. The curves for the different clay minerals are spread over a wide range. The value of such a comparison can readily be seen by considering the case of a soil dominated by kaolinitic clay such as the Cecil in comparison with a beidellitic soil such as the Putnam. Let us suppose that both are 10 per cent saturated with potas- sium and 90 per cent with calcium. The ionic concentration ratio K/Ca is thus 0.2. Then in the kaolinite soil the activity ratio K/Ca is 0.8 while in the beidellitic soil it is 4.0. This means an enormous difference in the chemical environment of the plant root in the two cases. Thus a considerable variety of factors play a part in establishing the mean chemical environment of the plant root. The nature of the col- loids, their concentration, and the particular proportions of exchange ions present are all powerful in their effects. Much further investigation in this field is needed, but the experimental methods are well under- stood. They have proved reliable precisely in those ranges of activities in which plant roots live and grow. CATIONIC ACTIVITIES AND PLANT GROWTH From the composition of the nutrient solution used by plant physi- ologists it is possible to calculate approximate cationic activities. It has been shown that plants thrive over a considerable range of such activi- ties, the range being widest where the cations concerned (usually K, Mg, and Ca) are in proper balance. So far as the author is aware, no studies have ever been published in which colloidal systems and true solutions have been compared at equal cationic activities as regards their effects upon plant growth. Such experiments are now entirely feas- ible. E. O. McLean (6) has made a preliminary study of this kind using montmorillonite clay with calcium and potassium together, and com- paring this medium with dilute true solutions of the same activities. Soybeans were grown in both sets of systems at a prescribed series of calcium and potassium activities. In order to maintain constant condi- tions, frequent changes of the nutrient systems were made. The over- all comparison of the effects of the solutions as compared with the C. Edmund Marshall 77 colloidal suspensions was vitiated to some extent by the fact that the bentonite systems released some magnesium to the plants, whereas none was supplied in the true solutions. The plants grown on the calcium-potassium bentonite systems were uniformly larger than those grown on the true solutions but were of the same percentage composi- tion as regards potassium and calcium. The most surprising feature of these experiments was that the systems giving 0Ca = 0.7 X io-4 and aK = 1.8 X io-4 apparently provided ample supplies of both calcium and potassium, since larger amounts and varied ratios of calcium and potassium showed no increase in uptake. Future experiments will therefore have to be designed to give comparisons at even lower activi- ties. The experimental methods for measurement will therefore be pushed to the utmost, since our membranes begin to deviate from the ideal behavior according to the Nernst equation below io-4 molar in the case of potassium and below io— 5 molar in the case of calcium. We intend to repeat this type of experiment with better control of the colloidal medium since it is of such critical importance. REFERENCES 1. Barber, S. A., Ph.D. Thesis, Univ. of Missouri (1949). 2. Burd, J. S., Soil Sci., 64:223 (1947). 3. Chatterjee, B., and Marshall, C. E., /. Phys. Chem., 54:671 (1950). 4. Jarusov, S. S., Soil Sci., 43:285 (1937). 5. Jenny, H., and Ayers, A. D., Soil Sci. 48:443 (1939). 6. McLean, E. O., Ph.D. Thesis, Univ. of Missouri (1948). 7. , and Marshall, C. E., Soil Sci. Soc. Am. Proc, 13:179 (1948). 8. Marshall, C. E., Soil Sci. Soc. Am. Proc, 7:182 (1942). 9. , and Ayers, A. D., Soil Sci. Soc. Am. Proc, 11:171 (1946). 10. , /. Am. Chem. Soc, 70:1297 (1948). 11. Marshall, C. E., and Bergman, W. E., /. Am. Chem. Soc, 63:1911 (1941). 12. , /. Phys. Chem., 46:52 (1942). 13. , /. Phys. Chem., 46:327 (1942). 14. Marshall, C. E., and Eime, E. O.. /. Am. Chem. Soc, 70:1302 (1948). 15. Marshall, C. E., and Krinbill, C. A., /. Phys. Chem., 46:1077 (1942). 16. Marshall, C. E., and McLean, E. O., Soil Sci. Soc. Am. Proc, 12:172 (i947)- CHAPTER 4 The Availability of Soil Anions ROY OVERSTREET AND I. A. DEAN T he availability of soil anions might well be considered as the state of being sufficient for the use of plants. Such a definition cannot readily be given a quantitative interpretation. Mineral nutrient availability usually embraces an integration of factors influencing the absorption of ions from soils and most studies concerned with availa- bility give consideration to the factors which influence it. For example, organic matter and liming are believed to increase while fixation and lack of movement of phosphate ions tend to reduce phosphate availa- bility. Indices of availability vary widely. Measurements such as the total amounts absorbed during the crop season and rates of absorption over a short period of time have been used. Nitrates, sulfates, phosphates, borates, and molybdates are soil anions essential for plant growth. In addition, plant ash contains chlorides and silicates in important quantities. Available anions exist either in the soil solution or in a solid phase in equilibrium with this solution. This system contains — in addition to the anions absorbed by plants — carbon- ate, bicarbonate, hydroxyl, fluoride, humate, and sometimes arsenate and arsenite anions which frequently have an important bearing on the state of equilibrium of a soil water system. By and large, consideration of anion availability has been restricted to a given plant nutrient. This discussion is an attempt to bring together the factors involved with the availability of soil anions in general. ORIGIN AND DISTRIBUTION OF SOIL ANIONS The essential elements absorbed by plants in the form of anions exist in soils in several discernible forms. These will be discussed under several different headings. 80 Mineral Nutrition of Plants Primary minerals Phosphorus, sulfur, boron, and molybdenum are components ot many primary minerals. These elements may constitute a definite part of the crystal lattice, as in the case of apatite, pyrites, or tourmaline, or they may be occlusions or isomorphic substitutions. From the stand- point of an immediate source of anions for plant growth, these minerals are of little interest. Minerals such as apatite and tourmaline (4) when ground to pass a 100-mesh sieve do not supply phosphorus or boron at a rate sufficient for the normal growth of most plants when supplied in cultures in amounts comparable with which these elements usually are found in soils. Organic compounds Probably one of the most significant differences between the availa- bility of soil anions and soil cations is that nitrates, sulfates, and phos- phates are produced in soils as a result of biological activity. Also, the state of oxidation of nitrogen and sulfur is altered by biological activity. It is generally assumed that most of the organic nitrogen in soils is protein in nature. The protein character of soil nitrogen has been variously studied. The carbon-nitrogen ratio is frequently used as an index of the nitrogen status of soils. This ratio (varies between 8 and 20) is a function of specific environmental conditions. The rate of nitrate formation in soils usually correlates more closely with the total nitrogen supply (/) than with measurements of the quality of the nitrogen compounds present in the humus. A relatively large proportion of the total soil phosphorus frequently is associated with the soil organic matter. For example, Dean (to) measured the organic phosphorus content of 34 surface soils from widely separated parts of the world. The organic phosphorus content of these soils varied from 8 to 50 per cent of the total phosphorus and the carbon-organic phosphorus ratio from 44 to 160. The identity of the organic phosphorus compounds in soils is somewhat obscure. How- ever, it is commonly believed that nucleic acids phytin and their deriva- tives (7, 49) account for a considerable part of the soil organic phos- phorus. Incubation studies (46) have indicated that important amounts of the organic phosphorus are mineralized. Over street and Dean 81 Evans and Rost (/6) have shown the organic sulfur in many Min- nesota soils to be an important part of the total soil sulfur. From 16 to 79 per cent of the total sulfur in these soils was found to be organic. The carbon-organic sulfur ratio for Minnesota soils varied from 87 to 451. There have been no studies on the specific nature of the organic sulfur compounds in soils. Little is known about the mineralization of soil sulfur in relation to the total supply of available sulfate. Anions associated with the clay fraction of soils Phosphates, boron, and molybdenum tend to accumulate in soils in association with the clay fraction. Analyses of clay separated from soils quite generally show a large part of the soil phosphorus and boron to be associated with this fraction. Analyses of soils have shown heavy ones to contain greater quantities of boron and molybdenum than those of light texture. What is known about the chemical nature of the anions associated with clays has been arrived at by indirect methods. Mineralogical and physical methods for the examination of clays have failed to show the presence of crystalline compounds of phosphorus or boron. Only through studies of the reactions between clays and solutions containing phosphates or borates — that is, phos- phorus or boron fixation studies — is it possible to arrive at many of the properties of the phosphates and borates associated with the clay frac- tion of soils. The availability of the anions associated with clays has not been satisfactorily understood. The equilibrium between the clay surfaces and the soil solution is considered by many to be a dominant factor controlling the availability of phosphates in soils. It is not improbable that the availability of borate and molybdate ions is similarly con- trolled. Soil solution and soluble salts Studies of the water-soluble material in soils date back to very early attempts to relate soil composition with nutrient uptake by plants. The inadequacy of the soil solution theory is probably the key to our present interest in nutrient availability. The amounts of soluble ma- terials in soils range from very small amounts in the acid soils of the humid regions to the saline conditions of soils in the arid regions. The 82 Mineral Nutrition of Plants soluble materials in soils have been considered in detail by Reitemeier (41), Parker (40), Anderson (2), and others. It must be concluded that the soil solution for a given soil is highly dynamic. The role of the soil solution in the absorption of anions by plants has never been wholly clarified. The amount of phosphorus, boron, and molybdenum present in the soil solution at any given time is inadequate for the nutrition of plants. In order that sufficient of these nutrients be available, it is necessary for one of two conditions to pre- vail; namely, that the soil solution be continuously renewed or that plants obtain these ions directly from the solid phase by contact feeding. Studies by Dean and Rubins (//) tend to minimize the importance of contact feeding as a means by which plants can obtain anions from soil surfaces. Studies by Volk (50) and Hunter and Kelley (22) on the absorption of ions from dry soils indicate that phosphates are not absorbed by plant roots from dry soils, whereas cations are absorbed. Studies on the effect of irrigation on the absorption of phosphates by sugar beets have shown absorption to be correlated with soil moisture. Judging from the rather scanty information available, it is not im- probable that plants absorb anions from soils through the medium of the soil solution. FACTORS INFLUENCING ANION AVAILABILITY Soil organic matter and biological activity Aside from being a possible source of available phosphorus, the organic matter itself is credited with having the capacity of increasing the availability of the soil phosphorus. Other reports show that humus reacts with rock phosphate, making it more available. Possible mecha- nisms by which humus makes inorganic soil phosphorus more available are not clear. Experimentally it has been shown by Steele (43) that less phosphorus is adsorbed by clays when humates are present in the system. An interesting series of experiments was reported by Gerretsen (17) who showed that more phosphorus is absorbed from insoluble phos- phates when in the presence of microbiological activity than from sterile cultures. Over street and Dean 83 Soil texture Soil texture can be shown to have various effects on availability. It is the common practice of European workers in this field (75) to take into consideration soil texture when evaluating soils for available nutrients. In general, light-textured soils need to contain higher amounts of readily soluble phosphorus than do heavy soils in order to have equal available phosphorus. At saturation, a unit volume of a heavy soil contains more soil solution than a light soil. The effect of this factor often has not been taken into consideration. Other things being equal, heavy soils fix more phosphorus. In so far as leaching or lack of leaching affects the amount of leaching and movement of ions in a soil, its texture influences availability. Change in pH and lime status A change in soil pH brings about changes in the equilibrium be- tween the soil solution and the solid phase of soils. The following experiment by Dean and Rubins (//) illustrates this point well. Six samples of soil colloid were suspended in 200 ml. of water and con- tinuously agitated. Varying amounts of monobasic sodium orthophos- phate were added to the suspensions representing approximately 5, 10, 15, 20, 25, and 30 per cent of the saturation capacity. Collodion bags were introduced into each system, and at intervals small aliquots were removed for analysis. Twenty-eight days after the start of the experi- ment an equal amount of sodium hydroxide was added to each of the suspensions. In Table I are given the phosphorus concentration and pH values of the intermicellar liquids. These data show the magnitude of change of concentration with pH. Equilibrium constants for the sets of data are relatively close. The effect of change in pH on the phosphorus, boron, and sulfate fixation will be discussed in a later section. Liming of acid soils is reported to increase the availability of the soil phosphorus. Salter and Barnes (42) in summarizing the long time field experiments at Ohio concluded "These facts are interpreted as indicating an increase in the availability of native soil phosphorus as the reaction is made more alkaline up to about pH 7.5." 84 Mineral Nutrition of Plants At one time it was believed that liming decreased the boron availa- bility of soils. Subsequent work, however, indicates that if it does affect availability, it is because of the relation to boron fixation by soils (5). Other investigations have shown that boron fixed by soils under alka- line soil conditions is readily released by acidification. TABLE I Distribution of Phosphorus Between Clay Surfaces and the Liquid Phase of a Clay-Water System as a Function of pH, and Amounts of Phosphorus Added* P Ap- Composition of Intermicellai • Liquid plied as Approx. Per- 22 di lys 35 d< iysf centage Satu- P pH Kt p pH KJ ration p.p.m. p.p.m. 5 0 002 51 0.85 X 10 3 0.059 6.9 1.6 X 10 a 10 0 006 5-2 0.74 X io~3 0.069 6.7 i-9 X 10 3 •5 0 01 1 5-3 0.80 X 10 0.079 6.6 2. 1 X 10 3 X io~3 20 0 .026 5-7 1.2 X 10 °-345 6.6 0.61 25 0 030 5-6 I.I X 10 0.262 6.8 1.8 X 10 3 3° 0 077 5-9 i .0 X 10 o-533 6.5 0.51 X 10 3 *Adapted from Dean and Rubins (//). fSodium hydroxide was added on the 28th day. _ Cqh (solution) X ChsPo, (surface) Coh (surface) X Ch3poi (solution) Robinson* has shown that crops grown on limed plots contained a greater concentration of molybdenum than the crops from adjacent unlimed plots. Soluble salts The concentration of phosphorus in the liquid phase of soils is dominantly influenced by the concentration and kind of salts present. The extent to which the concentration of phosphorus in the soil solu- tion is increased by lowering the salt concentration can be demon- strated by comparing the phosphatic concentration of successive in- *W. O. Robinson, private communication. Over street and Dean 85 crements of displaced soil solution. The effects hold for acid soils of the humid region as well as for calcareous and other soils of the arid regions. Mattson et al. (j2) have found neutral salts to increase the uptake of phosphorus by plants. Movement and leaching The movement of anions in soils is dependent upon the species of ion involved, the extremes being a comparison of the highly mobile nitrate ions with the much less mobile phosphate ions. Losses of ions through leaching may be reasonably considered as a factor influencing availability. The upward movement of ions in soils should also be con- sidered. The very restricted movement of phosphate ions in soils may be a restriction on availability. This becomes increasingly apparent if the distribution of phosphorus is thought of as a mosaic throughout the soil mass. With a lack of migration of ions the effectiveness of a given root system would be reduced. THE RETENTION OF ANIONS BY SOIL From the foregoing discussion on the factors influencing the availa- bility of soil anions, it may be seen that these factors are all in some way connected with the reactions that take place at the liquid-solid interface of soils. This section will deal with the kinds and mechanisms of these reactions. Retention of phosphate ions by soil The retention of phosphate by soils can be readily demonstrated by introducing soil into a solution containing phosphate ions and noting the decrease in concentration. This phenomenon is frequently termed "phosphate fixation." A review of this subject is available elsewhere (/j). The mechanism of phosphorus retention has not been adequately established. Apparently no single mechanism is applicable to all soil conditions. The older theories of phosphate retention by soils envision chemical precipitations. In acid soil systems iron and aluminum appear to be most likely agents, while in neutral or alkaline soils calcium and other 86 Mineral Nutrition of Plants divalent bases prevail. In acid soil systems solubility and plant growth studies point to possible inadequacy of a precipitation theory. In cal- careous soil systems, however, a precipitation theory seems adequate. Since the amount of phosphorus taken up by soils is proportional to the concentration, phosphorus retention has been ascribed to adsorption reactions. Adsorption is the tendency to concentrate at the interface. The term is general and may include several kinds of surface reactions. It does not carry any implication relative to the binding forces. Obser- vations have shown that some of the phosphorus retained by soils is more tightly held than others. For example Mattson and Karlsson (]i) have distinguished between colloid-bound phosphorus, a nondiffusable structural unit, and saloid-bound phosphorus, a diffusable ionic atmos- phere held as compensation to ions of opposite charge. The substitution of one ion for another or a metathetical reaction involving chemical forces and affinities is a means by which anions in solution may become associated with the solid phase or adsorbed. Thus, phosphate retention may be considered as the exchange of H2P04 ions in solution for OH ions associated with the solid phase. An increase in pH of the system is associated with this reaction; conversely, phosphates of soils may be displaced by hydroxyl, fluoride, or arsenate ions (12, Isotopic exchange studies (jo) between P3204 ions introduced into the soil solution and P3104 ions associated with the solid phase have indicated that only a small proportion of the phosphate ions retained by soils are readily exchangeable with similar ions in the liquid phase, whereas similar studies (6) have shown that the exchangeable calcium ions were all in equilibrium with the calcium ions in the liquid phase. The phosphorus retention by soils, for the most part, is restricted to the clay fraction. However, a large number of minerals common to soil exhibit, when finely ground, a capacity to fix phosphate. In acid soils particular attention has been given to the hydrous oxides of iron and aluminum and to the clay minerals. Maximum retention by these materials usually occurs in the pH range of 4 to 6. Soils treated to remove the iron oxide invariably show a reduction in their capacity to fix phosphorus (47). Overstreet and Dean 87 Retention of chloride and sulfate ions by soil The retention of chloride and sulfate ions by soil is not as readily apparent as the phosphate retention. The mechanism by which these ions are retained in acid soil systems in some respects parallels that for phosphate. Toth (48) measured the adsorption of chloride by clays before and after the free iron oxides had been removed (see Table II), and found TABLE II Adsorption of Chloride Ion by Untreated and by Deferrated Colloids* cr added as HCl-f NH4OH Free Iron Oxide Content pHof natant Super- Liquid CI Adsorbed Colloid Original Defer- rated A Original Defer- rated A Original Defer- rated A Cecil m.e./io g 20.00 20.00 20.00 per cent 12.32 1.44 3-7° 2.00 1.80 3-7° 2. 10 1.80 m.e./io g. 0.320 0.000 0.480 0.000 0.520 0.020 *Adapted from Toth (48) the untreated clays to adsorb this ion. The decrease in chloride adsorp- tion with increasing pH indicates the possibility that under acid con- ditions CI ions replace OH ions associated with the free iron oxides. Barbier and Chabannes (j) studied the retention of sulfate ions in soils and reached the following conclusions. 1. Sulfate ions are retained by soils more strongly than Cl ions but less strongly than P04 ions. 2. Calcium ion favored the retention of S04 ion independently of the precipitation of CaS04. 3. Soils of average composition contain 10-20 mg. of sulfur per kilogram in the adsorbed condition. Reitemeier (41), Kelley (28), and others have suggested sulfate adsorp- tion to explain the increases in soluble sulfate on dilution of some soils. 88 Mineral Nutrition of Plants Retention of boron by soils Apparently the mechanism for retention of boron by soils does not parallel that of the other soil anions (CI, S04, P04). Boron retention is lowest in acid soils, but increases rapidly in the range pH 6-10. Olson and Berger (^5) have shown that fixation is not affected by calcium addition except when this addition influences the soil pH. Boron fixa- tion by soils is readily reversible. Clays which have had the free iron and aluminum oxides removed from them fix more boron than un- treated samples. Negative anion adsorption There is accumulating evidence that in neutral and alkaline soils there is negative adsorption of nitrate and chloride ions. Reitemeier (41) and others have observed decreases in soluble nitrate and chloride in soils on dilution. In other words, there are lower concentrations of these ions in the solution immediately surrounding soil particles than at some distance from them. It has been suggested that this phenomenon may be caused by the insolubility of nitrate and chloride in the "unfree" or bound water. Another concept of the mechanism of this negative ad- sorption is that it is caused by a diffused anion swarm, the distribution of which depends on the Donnan distribution principle. ABSORPTION RATES AS A MEASURE OF ION AVAILABILITY A widespread experimental method for the study of ion availability is the measurement of absorption rates with selected test plants, or plant tissues under varying conditions in the soil or culture medium. While providing a facile means for investigation, this approach, in order to be productive, requires an adequate comprehension of the absorption process in plants and a knowledge of the physical chemistry of ions in soil. Unfortunately, this essential knowledge is not available to us, although a number of pertinent facts begin to appear. It will be our purpose in the following sections to list the established facts con- cerning ion absorption in plants as well as certain physical-chemical observations regarding ions in soil. Particular attention will be paid to anions, although the majority of the considerations apply to cations as well. Over street and Dean 89 Salient features of the ion absorption process in plants No mechanism proposed thus far for the accumulation of ions by plant cells has received universal acceptance. Moreover, it will not be our purpose here to describe in detail the various mechanisms that have been put forth nor to judge them. On the other hand, we wish to de- scribe a body of uncontested observations concerning the process which has been contributed to by a number of groups during the past 20 years. These groups include those associated with S. C. Brooks, R. Collender, D. R. Hoagland, H. Lundegardh, W. J. V. Osterhout, R. N. Robertson, and F. C. Steward. This information can be outlined as follows. 1. The ion absorption process requires the expenditure of energy by the plant (isotopic exchanges excepted). No ion accumulation occurs in the absence of respiration and other metabolic activities such as protein syn- thesis, etc. In general, the process is an attribute of tissues capable of growth. When the metabolic activity of the plant is inhibited by reduced oxygen tension, lowered temperature, or poisons, ion accumulation is like- wise inhibited. This is true for the accumulation of both anions and cations. 2. The ion absorption process is an exchange process. Predominantly cations are absorbed in exchange for H-ions of the plant and are released to the culture medium. Anions are absorbed in exchange for OH" or HC03~ which are released to the culture medium. The evidence indicates that no ion passes in or out of a healthy plant except by exchange for another ion. 3. Ion accumulation is to a large extent selective. Due to the exchange character of the process, anions can enter the plant independently of cations and vice versa. Also, ions are not absorbed at the same rates. In general, the cations K+, NH4+, Rb+, and Cs+ are rapidly accumulated while Ca , Mg+ , and Ba++ are much more slowly taken up. The anions N03 , Br-, and Cl~ are usually rapidly absorbed. The anions S04 and H2P04~ are moving rather slowly; the anion HC03~ is apparently not absorbed at all. 4. With electrolyte solutions, and below concentrations of the order of 0.005 N, the rate of accumulation of an ion is dependent on its concen- tration in the culture medium. The above-listed facts were essentially all that were firmly established concerning the ion absorption process before the advent of artificially prepared radioactive elements. The absorption process was known to be an exchange reaction which could be written formally as follows. 90 Mineral Nutrition of Plants For cations, R • H + K+ > R • K + H+ and for anions, R' • OH + Or > R' • CI + OH_ where R and R' symbolize the plant root. It was not known whether the combinations R-K and R'Cl represented a chemical binding or some physical entrapment. With the introduction of the use of radio- active isotopes, certain important advances have resulted in our com- prehension of the fixation of ions by plants. These advances have largely resulted from the study of isotopic exchange reactions. The significance of isotopic exchange reactions in the study of the absorption process The study of isotopic exchange reactions has been employed exten- sively by chemists to determine the nature of chemical binding. The most general expression for an exchange reaction may be expressed by the following equation, where A* represents the tagged or radioactive isotope and A represents the untagged or stable isotope: A*Y + AX^=±AY+ A*X Starting with A*Y, measurements of the rate of exchange can be accomplished by noting the appearance of A*X on the right-hand side of the equation. This rate is indicative of the type of chemical combi- nation between A* and Y. In general, if A* and Y are held together by electron pair bonds, the exchange reaction does not proceed unless some side reaction is possible that involves a breakage or elimination of the covalent bond. If the bond between A* and Y is electrovalent in character, the reaction ordinarily proceeds very rapidly, provided the ions entering into the reaction are physically accessible for exchange. Isotopic exchange reactions were used for the first time to study the nature of the combination of inorganic elements with plant roots by Hevesy in 192^ (18). Hevesy studied the rate of the reaction R • Pb* + Pb++ > R • Pb + Pb*++ where R symbolizes the plant root. Hevesy found that with Vicia faba plants, complete isotopic equilibrium was attained in 24 hours between 0 ver street an d Dean 9 1 absorbed lead and lead in the culture solution. From this fact he con- cluded that lead is held in the plant in the form of a dissociable saline compound, that is, by electrovalent forces. At a much later date (19,20,21) Hevesy and his co-workers carried out similar experiments with radioactive phosphate. They concluded that in the case of a number of plant parts the phosphorus was held by electrovalent forces, presumably as inorganic phosphate, although in the case of a yeast sample it was probably held by electron pair bonds — possibly as hexosephosphate or adenylphosphoric acid. In 1938 Jenny and Overstreet (24) studied the rate of the reaction R • K* + K+ > R . K + K*+ using potassium clays and potassium chloride solutions for the outside media. They found that although the reaction proceeded at a measur- able rate, only about 10 per cent of the K* initially held by the roots underwent exchange in a period of 24 hours. These studies were soon followed by similar experiments with radio- active sodium, rubidium, potassium, bromine, and phosphate by Jenny, Overstreet, and Ayres (26), Mullens and Brooks (jj), Overstreet and Broyer (jS), and Broyer and Overstreet (9). In brief, the experiments showed that all the above-mentioned elements are held by the plant in an exchangeable form, and therefore most likely not entirely by elec- tron pair bonds. On the other hand, a great diversity was noted in the rates by which individual elements underwent exchange. This indicated considerable variation in the strength of binding of the different ele- ments in the plant. Recently, the exchange of ions at extremely low concentrations be- tween barley roots and culture solutions has been studied by Overstreet and Jacobson (39) and Jacobson and Overstreet (23). The exchange behavior of Rb+, HL.P04— , Sr++, and I- at concentrations of the order of io-9 molal was studied. The advantages of such solutions are (a) freedom from osmotic effects and (b) extreme degree of sensitivity, enabling experiments to be carried out with single root tips for as short a time as 30 seconds. The isotopic exchange curves on individual apical segments were determined at temperatures near o° C. in order to avoid complications due to metabolic activity. The experiments showed that the isotopic exchange curves for the 92 Mineral Nutrition of Plants various ions were characteristically different. This fact will be apparent from Figure i, which gives the curves for Rb+, Sr++, FLP04~, and I- . Also, it was found that in each case the isotopic exchange curve was characteristic of the living root. With ether-killed tissue, all the ele- ments were released very rapidly by isotopic exchange. This point is illustrated in Figure 2 for the case of I~. 50- I I ^ 5 (J"-^--. "H2P04 \ 51 ^s""®"> T 3 f\ T - ^^ ^""""■—KSX-^. ■ ""IS)***. — — Sr _^_Rb+ , 1 , , i , 1 i 10 TIME IN MINUTES 15 20 Figure i. Graphs showing the release at O0 C. of P32, I131, Srs5, and RbS6 by apical root segments in exchange for the inactive isotopes in the culture solution. Apical segments of roots (2 cm.) containing the absorbed radioelements were placed in 0.005 N solutions of H2P04~, I-, Sr++, and Rb+, and counted at selected intervals to determine the loss of the radioactive elements by exchange. The percentage of the initially absorbed activity remaining in the root is plotted against time.— Overstreet and Jacobson (38), and Jacobson and Overstreet (23). A third point of interest was the fact that, at least in the neighbor- hood of o° C, the different elements showed rather widely different longitudinal absorption patterns in the root tips. That is, for example, regions in the tissue which absorb H2P04- very rapidly do not neces- sarily absorb I- rapidly. This finding is illustrated in Figure 3. In general, the isotopic exchange experiments seem to indicate that Over street and Dean 93 the combinations RK, RNa, R'Cl, R'l, R'H,P04, etc., represent chemical attachments. It would be extremely difficult to interpret the two-way movement of ions characteristic of the isotopic exchange re- actions on the basis of a mere physical entrapment. It must be concluded further in the light of the exchange experi- ments that the various combinations represent either different chemi- Live Roots TIME IN MINUTES Figure 2. Graphs showing the exchange for inactive isotopes of I131 at O0 C. for live and ether-killed roots. The percentage of the initially ab- sorbed activity remaining in the root is plotted against time. — Jacobson and Overstreet (23). cal entities or at least similar complexes with widely varying energies of formation. Finally, it must be concluded that the binding complexes are rather labile, since the mere injury or killing of the tissues is sufficient to de- stroy them. The idea that ion absorption is chemical in nature is not in conflict with a number of absorption theories that have been put forth — some of long standing. Some of these postulations are of particular interest here. 94 Mineral Nutrition of Plants 3.00 5 §2.00 6 K1.00 I 3 4 5 6 7 MM. FROM ROOT APEX 8 10 ri3i Figure 3. Graphs showing the distribution of absorbed P32 and Iloi in barley roots as a function of distance from the root apex. The dotted line labeled "external solution" corresponds to the activity of a volume of the bathing solution equal to that of 1 mm. of root segment and is arbitrarily given the value of 1.00 in the graph. All other amounts are given in terms of the value of the external solution. — Overstreet and Jacobson {38), and Jacobson and Overstreet (23). Hypotheses concerning the chemical nature of R-H and R'-OH On the basis of certain investigations dealing with the mechanism of ion accumulation by plant cells, Osterhout (j6) postulated that the plant substance R-H may be similar in its properties to some of the aromatic alcohols. Osterhout was successful in constructing an artificial model of a cell which consisted of two electrolyte solutions separated by a layer of potassium guaiacolate. By maintaining a difference in hydrogen ion concentration between the electrolyte solutions, he was able to show an accumulation of K-ions in the solution highest in H- ions. No hypotheses were made concerning the substance R' • OH. Over street and Dean 95 On the other hand, Brooks (8) became interested in the role of amino acids in plant cells. He was led to the theory that the properties of both of the substances RH and R'-OH were inherent in the amino acid molecule; the H+ of the -COOH being exchangeable for cations and the OH- of the -NH5OH groups being exchangeable for anions. By postulating an orientation of amino acid molecules in the protoplasm or plasma membrane (mosaic arrangement) Brooks was able to en- visage the entry into the cell of both cations and anions along different paths by a series of ion exchanges involving H+ and OH-. At a much later date Steward and Street (44), on the theory that ion accumulation is intimately tied up with protein synthesis in the plant, speculated that RH and R'-OH may correspond to the acidic and basic groups of certain phosphorylated energy-rich nitrogen compounds in the proto- plasm. Very interesting speculations regarding the nature of R-H and R'-OH are embodied in Lundegardh's theory of "anion respiration" (29). According to Lundegardh there is little difficulty in accounting for the properties of R-H in plants, since the protoplasm as a whole is negatively charged and contains appreciable quantities of substances with comparatively strong acid properties. Cations in the culture me- dium therefore exchange for H+ ions in the plasma membrane and proceed inward through the protoplasm by exchange along paths or tracks of substances of acid dissociation. No special energy would be required for this type of transport. For these reasons Lundegardh's theory is chiefly concerned with the nature of R'-OH, that is, with the means by which anions are absorbed and moved within the plant. Accordingly, Lundegardh conjectures that ". . . the Fe-ion in the hemin group of a respiratory enzyme is indeed well suited to effect an anion transport, according to the following scheme 'Fe+ + + \ +e /Fe++\ 3A~ / -e \2\~ / "The trivalent Fe-atom attracts one more anion than the bivalent Fe- atom. If the enzyme system constitutes a structural unit, in which elec- trons move from one atom to another in the next molecule . . . the 96 Mineral Nutrition of Plants conditions are given for a transport of anions. The change of valency proceeds as an electron wave, if the enzyme system lies parallel to a redox gradient. If the molecules of the enzyme are arranged in such a way that they serve as a boundary between two media of different redox potentials, owing to the wavelike proceeding oscillation of the Fe-valency, they will transport anions from the medium with higher oxydation power (Fe+++; ox-side) to the medium with lower oxyda- tion power (Fe++; red-side). The anions are transported in opposite direction to the electrons. . . . The transference of an electron between two Fe-atoms moves one anion from the oxidized to the reduced stage." On the basis of the character of the isotopic exchange curves, Jacob- son and Overstreet (2]) were led to conclude that the plant complexes R-K, RRb, RSr, R'H2P04, R'l, etc., probably do not represent combinations of the ordinary electrovalent type. They suggest that cations may be bound in plants in the form of chelated complexes. In this class of structure the metallic ion is attached at two or more points in the same molecule and one of the bonds is frequently coordinate in nature (see Yoe and Sarver, Organic Analytical Reagents, John Wiley and Sons, 1941). Many plant substances such as proteins, amino acids, and organic acids are known to form chelated compounds, particularly with polyvalent cations. Chelated structures in general are rather stable arrangements and metallic elements so bound do not readily undergo isotopic exchange. An example of this class of compound is given in the proposed structure for the calcium complex of ethylene-diamine- disodium-tetraacetate : NaOOCCH2 CH2-H2C CH2COONa PLC COO' The foregoing outline of hypotheses concerning the nature of the plant substances RH and R'-OH is by no means exhaustive. Neverthe- Over street and Dean 97 less, it is hoped that it gives a picture of the present state of uncertainty concerning these compounds and indicates a wide field for future re- searches. Information concerning the reversibility of absorption reactions As will be emphasized in a later section, a knowledge of the reversi- bility of ion absorption processes is essential for the formulation of soil factors affecting ion availability. Here again our information is very meager. In a study of the effects of suspensions of hydrogen clays on barley roots, Jenny and Overstreet (24) obtained results that perhaps serve as evidence for the reversibility of the reactions R • H + K+ > R • K + H+ and 2 R • H + Ca++ > R2 • Ca + 2H+ The plant roots were observed to lose significant fractions of their potassium and calcium contents after a few hours of contact with dilute hydrogen-bentonite suspensions. Evidently this loss was not the result of permanent injury, since subsequently it could be shown that the tissue absorbed normally from standard nutrient solutions. In so far as can be determined, evidence for the reversibility of anion absorption is even less substantial than in the case of cation absorption. In some unpublished work by J. M. Heslep, plant roots were placed in suspensions of clays of the kaolinite type. In a few hours significant amounts of phosphate were observed to leave the roots and become fixed on the clay. This may be taken as presumptive evidence for the reversibility of the reaction R' • OH + H2P04" > R' • H2P04 + OH". A series of experiments testing the reversibility of absorption re- actions in general are being undertaken in the Divisions of Soils and Plant Nutrition at Berkeley. The effects of concentrations and activities on absorption rates in nutrient solutions Ion-absorption experiments conducted with flowing nutrient solu- 98 Mineral Nutrition of Plants tions, or with nutrient solutions maintained at constant composition, definitely show that up to concentrations in the neighborhood of 0.001 N, the absorption rate of an ion is a linear function of its con- centration in the culture medium (24,37). Since, in nutrient solutions as ordinarily used, the activities of ions are not appreciably different from their molalities, it is not possible to discern in these cases whether the rates of absorption of the ions by plants are determined by the concentrations or by the activities of the ions. This point, while per- haps not of great importance in the study of plants growing in nutrient solutions, becomes a question of major concern in the estimation of ion availability from soil systems. This becomes evident when the physical chemistry of soil ions is considered. Some conclusions regarding the activities of soil ions The characteristic chemistry of soil ions becomes apparent when a soil suspension or gel in equilibrium with its filtrate or "soil solution" is considered. This approach has practical as well as theoretical impli- cations, since many soil fertility tests are based on the chemical com- position of the soil solution rather than of the whole soil. Such an equi- librium system can be represented as outlined in Figure 4. SOIL SOIL SOLUTION Figure 4. Diagram showing a soil suspension or gel in equi- librium with its filtrate or soil solution. Over street and Dean 99 Phase I contains soil particles and ions such as Ca++, Mg++, Na+, and K+ in the adsorbed state. Also it will contain the aforementioned cations and anions such as CI-, HCO..-, NO.,-, H2P04-, and S04~ in the free or unadsorbed state. Phase II is purely an electrolyte solution which contains no soil particles. We may now consider the conditions for equilibrium between Phase I and Phase II. According to a law of thermodynamics, the partial molal free energy (F) of any component is the same in all phases of a system of phases at equilibrium. That is, for example, (FKCi)i = (Fkci)ii at equilibrium. (t) This thermodynamic law is perfectly general in character and applies for the case of charged ions as well as for the case of uncharged com- ponents. For this reason we can also write (Fk+)i = (FkOii (Fci-)i = (Fcl-)„ Since we know that the concentrations of individual ions in our sys- tem may be quite different in the two phases, it should be borne in mind that the partial molal free energy of an ion in such a system may be a function of factors other than concentration such as electric poten- tial or interaction with surfaces.* In conformity with the original definition of activities, we can define the activity (a) of an individual ion in our system by means of the equation: F - F? = RT In a-, where F" represents the partial molal free energy of the ion i in an arbitrarily chosen standard state. Consequently, we may write *It should be noted that in this approach the partial molal free energy of an individual ion is identified with the "escaping tendency" of the ion. It is identical with the electrochemical potential, fT, of another treatment and nomenclature (for example, see The Physical Chemistry of Electrolytic Solutions by H. S. Harned and B. B. Owen, New York, Reinhold Publishing Company, 1943, page 315). ioo Mineral Nutrition of Plants (RT In aK+ + Fk+)i = (RT In aK+ -f Fk+)ii (RT In aGl- + FCi-)i = {RT In acl- -f FnGl-)u If we choose to express the activities in both phases in terms of the same standard state, we can write further («K+)l = («K+)ll (tfoi-)r = (tfci-)n Thus we are led to conclude that, on the basis of the same standard state, the activity of any ion in a soil suspension or gel is the same as in its equilibrium filtrate.* In some quarters it may be contended that this conclusion is contradicted by the well-known observation that the pH, as ordinarily determined with the hydrogen or glass electrode, may be quite different in suspensions of acid clays and in their filtrates. How- ever, it should be pointed out that such measurements invariably in- volve an electrolytic bridge and a liquid-liquid junction. The magni- tude of the junction potential cannot be measured. Conceivably, it may be quite different in a clay suspension and in a dilute salt solution. For these reasons, conclusions regarding H+ activities in systems of this kind based on pH determinations are subject to question. The activity of an ion can be expressed in terms of its concentration by means of the relationship * According to this treatment, the activity of an individual ion is defined in terms of its electrochemical potential. In a number of treatises (see, for example, J. N. Bronsted, Physical Chemistry, New York, The Chemical Publishing Com- pany, 1938, page 276) the electrochemical potential of an ion is considered as being composed of two parts, one chemical and the other electrical: i.e., u.; : Hi + z{ F i|», where JTj is the electrochemical potential, M-i the chemical potential, zx the valency of the ion, F Faraday's member, and •»!> is the electric potential of the phase. Moreover, some writers using this line of reasoning choose to define the activity of an individual ion in terms of its chemical potential, although they point out that the activity so defined has no thermodynamic significance. 0 ver street an d Dean i o I where m\ is the molality of the ion and Yi is its activity coefficient. Yi becomes unity in an infinitely dilute solution. Many soil filtrates are dilute to the extent that the activity coefficients of their constituent ions can be taken as unity and the activities of the ions can be considered equal to their molalities. In such cases, where the total concentrations in each phase are known, it is possible to calculate relative activity co- efficients for the ions in the soil suspension or gel. When this calculation is made, it is found very often that the activity coefficient of a soil cation is very much less than unity, often of the order of o.ooi. On the other hand, we find that with soil suspensions that display such Donnan effects as negative adsorption (41), the activ- ity coefficient of a soil anion may be considerably greater than unity. This, however, is not an uncommon situation even with ordinary elec- trolytes (cf. the mean activity coefficient of 2 M HCl). Ion absorption rates in soil suspensions and their filtrates In a series of studies dealing with the "contact" and "soil solution" theories, Jenny and Overstreet (24, 25, jj) were able to show that with certain clay or soil suspensions containing adsorbed Na+, K+, or Rb+, the rates of absorption of the ions by barley roots was greater than from the corresponding filtrates. In terms of the present argument, these ex- periments are rather strong evidence that the absorption rates of some soil ions are functions of their concentrations in the culture medium rather than their activities. That is, although the ion activities were the same in suspension and filtrate, the rate was greatest in the phase where the concentration was greatest. A similar situation has been encountered in other branches of chem- istry. Although most theories on reaction kinetics predict that rates will be functions of the activities of the reactants, a rather impressive num- ber of reactions have been discovered in which the rates are determined by concentrations {34). In our present state of knowledge, predications beforehand of the rate-determining factors of chemical reactions are uncertain. Also, we are probably not justified in making any generalizations concerning the absorption of soil ions. To date, the effect has been established for Na+, K+, and Rb+ ions onlv. 102 Mineral Nutrition of Plants Information as to whether activities or concentrations determine the absorption rates of anions is scanty indeed. Appropriate systems for testing this point would be soils that show large concentration differ- ences of anions between the soil suspension and its filtrate. These con- ditions presumably would obtain in soils which show a large negative absorption of anions or in soils that adsorb phosphate, arsenate, or molybdate in large amounts. The experiments of Dean and Rubin (//) with a soil high in fixed phosphate would seem to indicate that the absorption rate by plants of phosphate is a function of the activity of the phosphate ion rather than of its concentration. That is, the adsorption rate from a soil high in adsorbed phosphate was apparently no greater than from its filtrate. It is evident that a great deal more experimentation is warranted on this point. Nevertheless, the fact that the absorption rates of such im- portant soil ions as Na+ and K+ are controlled by concentration rather than activities gives rise to the question as to what functions of con- centrations in the soil are appropriate indications of the availability of these ions. Relationship of ion availability to certain concentration factors in the soil For many years attempts have been made to correlate the availa- bility of an ion with its concentration in the exchange complex of the soil. On the whole this correlation has not proved very satisfactory. Fairly recently, Jenny and Ayres (2j) and others have attempted to correlate the availability of Na+, K+, and Ca++ with their equivalent percentage in the exchange complex of the soil (degree of saturation). In a number of cases this procedure has resulted in fairly satisfactory correlations; however, as yet the results are so limited that generaliza- tions are not justified. In the case of anions that are not adsorbed by soil particles, it is customarily assumed that an adequate assessment of their availability is given by their concentrations in the soil water. Whether this assump- tion is justified in soils that exhibit a large negative adsorption of anions has not been determined. The availability of anions such as phosphate, arsenate, and molybdate that are adsorbed in large amounts by some O ver street an d Dean 1 03 soils is a field for a great deal more experimentation. In the case of phosphate it begins to appear that its availability may be rather closely related to its concentration in the soil solution; that is, in a filtrate of the soil. This is additional indirect evidence that the absorption rate of the phosphate ion may be determined by its activity rather than con- centration in the soil medium. Importance of the question of reversibility of the ion absorption reaction A knowledge of the reversibility of the absorption by roots of a given nutrient ion is of paramount importance in the estimation of its availability. This is the case because the effective rate at which a reversi- ble reaction proceeds will be influenced by the concentrations or activi- ties of the products as well as of the reactants. For example, if it were definitely established that the following absorption reaction were reversible R' ■ OH + H2P(V > R' ■ H2P04 + OH-, then obviously the extent to which H2P04_ would be absorbed would depend on the amounts of R'H2PO, in the plant and of OH- in the culture medium as well as on R'OH and H2P04~. Thus, eventually it may be established that the availability of certain ions are functions of ratios of concentrations or activities. REFERENCES 1. Allison, F. E., and Sterling, L. D., Soil Sci., 67:239 (1949). 2. Anderson, M. S., Keyes, M. G., and Cromer, G. W., U. S. Dept. Agr. Tech. Bull. 813 (1942). 3. Barbier, G., and Chabannes, J., Compt. rend., 218:519 (1944). 4. Berger, K. C, and Truog, E., /. Am. Soc. Agron., 32:297 (1940). 5. Berger, K. C, Adv. in Agronomy, 1:321 (1949). 6. Borland, J. W., and Reitemeier, R. F., Soil Sci. (In Press) (1950). 7. Bower, C. A., Soil Sci., 59:277 (1945). 8. Brooks, S. C, Trans. Faraday Soc, 33:1002 (1937). 9. Broyer, T. C, and Overstreet, R., Am. /. Botany, 27:425 (1940). 10. Dean, L. A., /. Agr. Sci., 28:234 (1938). 11. , and Rubins, E. J., Soil Sci., 59:437 (1945). 12. , Soil Sci., 63:377 (1947). 104 Mineral Nutrition of Plants 13. Dean, L. A., Adv. in Agronomy, 1:391 (1949). 14. Dickman, S. R., and Bray, R. H., Soil Sci., 52:263 (1941). 15. Egner, H., Kohler, G., and Nydahl, F., Ann. landw. Hochschule Schived., 6:253 (1938). 16. Evans, C. A., and Rost, C. O., Soil Sci., 59:125 (1945). 17. Gerretsen, F. C, Plant and Soil, 1:51 (1948). 18. Hevesy, G., Biochem. ]., 17:439 (1923). 19. , Linderstrom-Lang, K., and Olsen, C, Nature, 137:66 (1936). 20. , Nature, 139:149 (1937). 21. Hevesy, G., Linderstrom-Lang, K., and Nielsen, N., Nature, 140:725 (i937)- 22. Hunter, A. S., and Kelly, O. J., Plant Physiol., 21:445 (1946). 23. Jacobson, L., and Overstreet, R., Am. /. Botany, 34:415 (1947). 24. Jenny, H., and Overstreet, R., Proc. Nat. Acad. Sci. U. S., 24:384 (1938). 25. , and Overstreet, R., /. Phys. Chem., 43:1185 (1939). 26. , and Ayers, A. D., Soil Sci., 48:9 (1939). 27. Jenny, H., and Ayers, A. D., Soil Sci., 48:443 (1939). 28. Kelley, W. P., Soil Sci., 47:367 (1939). 29. Lundegardh, H., Ar\iv. f. Bot., 32A, No. 12:1 (1945). 30. McAuliffe, C. D., Hall, N. S., Dean, L. A., and Hendricks, S. B., Soil Sci. Soc. Am. Proc, 12:119 (1947). 31. Mattson, S., and Karlsson, N, Ann. Agr. Coll. Sweden, 6:109 (1938). 32. Mattson, S., Eriksson, E., Vahtras, K., and Williams, E. G., Ann. Roy. Agr. Coll. Sweden, 16:457 (1949). 33. Mullins, L. J., and Brooks, A. C, Science, 90:256 (1939). 34. Olson, A. R., and Simonson, T. R., /. Chem. Phys., 17:1167 (1949). 35. Olson, R. V., and Berger, K. C, Soil Sci. Soc. Am. Proc, 11:216 (1946). 36. Osterhout, W. J. V., Botan. Rev., 2:283 (1936). 37. Overstreet, R., and Jenny, H., Soil Sci. Soc. Am. Proc, 4:125 (1939). 38. Overstreet, R., and Broyer, T. C, Proc Nat. Acad. Sci. U. S., 26:16 (1940). 39. Overstreet, R., and Jacobson, L., Am. J. Botany, 33:107 (1946). 40. Parker, F. W., Soil Sci., 12:209 (1921). 41. Reitemeier, R. F., Soil Sci., 61:195 (1946). 42. Salter, R. M., and Barnes, E. E., Ohio Agr. Expt. Sta. Bull. 553 (i935)- 43. Steele, G. J., in Abstracts of Doctors' Dissertations, No. 15 (Columbus, O., Ohio State Univ. Press, 1935), p. 203. Overstreet and Dean 105 44. Steward, F. C, and Street, H. E., Ann. Rev. Biochem., 16:471 (i947)- 45. Stout, P. R., Soil Sci. Soc. Am. Proc, 4:177 (1939)- 46. Thompson, L. M., Soil Sci. Soc. Am. Proc, 12:323 (1948). 47. Toth, S. J., Soil Sci., 44:299 (1937). 48. , Soil Sci., 48:385 (1939). 49. Wrenshall, C. L., Dyer, W. J., and Smith, G. R., Set. Agr., 20:266 (1940). 50. Volk, G. M., /. Am. Soc. Agron., 39:93 (1947). CHAPTER w Contact Phenomena Be- tween Adsorbents and Their Significance in Plant Nutrition HANS JENNY S, peculations on the nature of mineral uptake by roots in soils are encountered very early in botanical literature. Two states of existence of nutrients in the soil were recognized: first, nutrients in the solid portion which were considered unavailable to plants; second, nutrients dissolved in the liquid phase which could be ready assimi- lated by plant roots. These dissolved nutrients were known to diffuse freely in the soil moisture, and they were known also to move with the flow of water in the soil. Water containing dissolved nutrients consti- tuted the soil solution, a fruitful concept which is still in use today. In its essence, the soil solution corresponds to the nutrient solution of the plant physiologist. SOIL SOLUTION CONCEPT Investigators who study the soil solution usually define it on an operational basis. The soil solution is that part of the liquid phase which can be separated from the bulk of the soil by some sort of displacement method (5), extraction or centrifugation. To the colloid chemist it is clear that the extraction or displacement technique does not measure the adsorbed (exchangeable) cations, in- cluding hydrogen, which remain attached to the colloidal particles. As we shall see, this omission has far-reaching consequences. In this paper we shall assign the adsorbed ions to the solid phase. Although they are surrounded by and bathed in water molecules, they 108 Mineral Nutrition of Plants are not free to diffuse in the liquid phase. The ion swarm always re- mains- attached to the colloid particle and goes wherever it goes. One might redefine the soil solution as comprising the dissolved electrolytes in the displaced soil extract plus the exchangeable cations and anions of the solid phase. Rather than introducing a new definition of soil solution, the writer prefers Marshall's term "ionic environment of the root" for the combination of adsorbed (exchangeable) ions and the classical soil solution ions. Probably the most extreme proponent of a simple soil solution theory of plant nutrition was Cameron (5). As late as 1911, he stated: "There can be no doubt, therefore, that the soil solution is normally of a con- centration amply sufficient to support ordinary crop plants, and is maintained at a sufficient concentration so far as mineral plant nutrients are concerned." To account for the existence of soils giving low yields, Cameron postulated and, seemingly, demonstrated that infertility of soils is caused by the presence of toxic or inhibitory organic substances in the soil solution. Cameron labored under the erroneous impression that the liquid phase was nothing but the saturated solution of sparingly soluble soil minerals. He overlooked the contribution of the adsorbed ions to the ionic environment of the plant root. Von Liebig, preceding Cameron by half a century, appears more farsighted. After some hesitation he repeated and confirmed Way's famous experiments on base exchange. In 1858 he expounded new ideas on plant nutrition (27) : "It is clear that we must abandon this idea (soil solution theory) when it can be demonstrated that rain water, either alone or in con- junction with carbonic acid, does not dissolve enough of the mineral constituents to contribute significantly to plant growth. In this case the uptake of minerals must be the result of an active contributing cause residing in the root, whereby the water surrounding the root is enabled to dissolve certain mineral constituents which, otherwise, it could not do. As further consequence, the quantity of mineral sub- stances consumed must be in proportion to the root surface of the plant and to the sum of those active mineral constituents that are con- tained in such portions of the soil as are in contact with the root sur- face." Hans Jenny 109 Loosjes (28) hinted that Liebig foresaw the modern contact theory of plant nutrition in soils. This contention is hardly justified. In Liebig's days two essential prerequisites of the contact exchange theory were unknown; namely, the law of mass action, especially in relation to its reversibility aspects, and the concept of the electric double layer. More- over, Liebig himself stated (p. 139) : "It is very difficult to visualize in what manner plants contribute to the solubility of mineral constituents; that water is essential to their passage is obvious." The idea that the extracted soil solution does not provide the whole answer to the question of mineral nutrition of plants in soils was sporadically advanced by several investigators, such as Comber (7), Kossovitch (25), and Truog (44). These attempts to attribute a more active role to the solid phase were unsuccessful for two main reasons. First, no reliable experimental material was produced which would render untenable the soil solution theory, especially if combined with the picture of carbon dioxide excretions. Second, and most important of all, no theory was advanced which provided a convincing reaction mechanism involving the solid phase. SALIENT ASPECTS OF THE CARBON DIOXIDE THEORY OF MINERAL NUTRITION OF PLANTS According to prevailing ideas, roots secure cations adsorbed on soil particles by means of carbonic acid exchange (Figure 1). The transfer involves the following steps: 1. Release of carbon dioxide from the root and formation of carbonic acid. 2. Diffusion of carbonic acid to the distant clay surface. 3. At the clay surface H+ replaces K+; the clay particle becomes acid. 4. The new ion pair, K+HC03"~ , returns to the root sur- face. 5. At the root surface K+ exchanges for H+, or K+HC03— enters the root as an ion pair. Of the last two alternatives the latter seems less probable. Overstreet, Ruben, and Broyer (j6) immersed barley roots in a solution of potas- no Mineral Nutrition of Plants sium acid carbonate in which the anion HCOa_ contained radioactive carbon (C11). Although the roots accumulated large quantities of potassium they utilized only a small fraction, 4-5 per cent, of HC03~. Jenny and Cowan (17, 18) grew soybean seedlings on pure calcium clay suspensions having an initial pH of 6.30. At harvest time the plants had removed from the clay 1.020 milliequivalents (m.e.) of calcium. The reaction of the sol dropped to pH 4.32 and the clay particles now contained 0.948 m.e. of hydrogen which they did not Root t~C02+H20-+H+HC0~3 - Na K+HCO\ Figure i. Conventional model of liberation of ad- sorbed ions by roots (C02-theory). possess at the beginning of the experiment. While this observation is in harmony with step 3 of the carbon dioxide theory, it is also in accord with the contact theory. In this connection an experiment by Overstreet, Broyer, Isaacs, and Delwiche (_?_?) is illuminating. These authors determined the uptake of cations and anions by barley roots from potassium solutions and potassium clay suspensions. They compared it with carbon dioxide evolution and the synthesis of organic acids in the cell sap of the root. The authors concluded "the excess accumulation of cations over anions is roughly balanced by organic acid anions (other than HCO:{) which are synthesized within the plant. Moreover, it is apparent that these synthesized organic acids are the ultimate sources of the hydrogen which replaces the adsorbed potassium on the clay, and not carbonic acid." INADEQUACIES OF THE SOIL SOLUTION THEORIES OF PLANT NUTRITION In discussing his experiments on plant growth in synthetic ion ex- change media, Arnon and Grossenbacher (j) concluded ". . . the data Hans Jenny in do not require the invoking of a contact exchange mechanism to ex- plain the results obtained." Allaway (2) expressed a similar opinion. Indeed, why should anyone wish to consider a new theory so long as the old one appears satisfactory ? A body of data exists which casts doubt on the omnipotence of the soil solution theory. We shall briefly discuss some of the more perplex- ing situations. Uptake of ions by living roots from salt solutions and clay suspensions For a period of 10-20 hours, excised, low-salt barley roots rapidly accumulate ions. It is of interest to compare the uptake of cations by roots from salt solutions and clay suspensions having equal concen- trations of cations. According to Figure 2, the uptake of radioactive sodium (Na#), at higher concentrations, is decidedly greater in clay suspensions than in chloride or bicarbonate solutions (20). This obser- vation is confirmed with nonradioactive sodium (^5). In the case of ammonium clay versus ammonium chloride, the uptake of ammonium by the roots is nearly the same for the two systems. On the other hand, potassium chloride provides a better source of potassium than potas- sium clay. These last two experiments were conducted by Ayers (4) who used the Hoagland technique (12). On the generally held assumption that ion uptake is a function of activities, only the potassium systems behave somewhat according to expectations. With the aid of clay membranes, Marshall (jo) has shown that salt solutions possess higher activities than clay suspensions having concentrations of cations equal to the salt solutions. Accordingly, salt solutions should provide more efficient nutrient media. It is not known how far the experiments with barley roots can be generalized. Suffice it to emphasize that the uptake of cations from clay suspensions bears no simple relationship to ion accumulation from salt solutions having corresponding cation contents. Significance of type of clay mineral in nutrient uptake by plants Several years ago Elgabaly, Jenny, and Overstreet (8) observed that under comparable conditions barley roots accumulate more potas- ii2 Mineral Nutrition of Plants sium and zinc from montmorillonite clay suspensions than from kaolinite clay suspensions. These observations are in striking contrast to what one would expect to happen on the basis of the carbon dioxide theory. Release of potassium and zinc from the two clay minerals by water, carbonic acid, and hydrochloric acid is greater for kaolinite than for montmorillonite. The soil solution or, more precisely, the supernatant liquid of the kaolinite suspension always contains more bases than the montmorillonite sol. Accordingly, nutrient uptake should be favored by kaolinite and not by montmorillonite as is actually observed. Experiments performed by Mehlich and Col well (?2) and by Allaway (2) who grew cotton and soy bean plants for several weeks produced results opposite to those of Elgabaly and his co-workers. Recently, how- ever, Elgabaly and Wicklander (9) confirmed the previous findings, showing that barley roots accumulate more calcium and sodium from montmorillonite systems than from kaolinite systems. We must con- clude that the experiments of Elgabaly et al. also are valid and that for their systems and techniques the soil solution theory — with or with- out the carbon dioxide mechanism — is incapable of explaining the results. Contact depletion Living barley roots low in potassium tenaciously hold onto their content of potassium. Even intensive leaching of the roots with 380 liters of distilled water will not remove measurable amounts of potas- sium (21). Likewise, electrolyte solutions comprising sodium chloride, hydrochloric acid, sodium bicarbonate, and ammonium bicarbonate will not significantly affect the potassium status of the roots. On the other hand, as will be discussed on page 121, hydrogen clay, calcium- hydrogen clay, sodium clay, and ammonium clay very significantly deplete the root system of potassium. More sensitive techniques involving radioactive potassium reveal similar root behavior. While some of the radioactive isotope is released by the plant to water and electrolyte solution, the outgo becomes es- pecially pronounced when the roots are in contact with colloidal clay particles. It should be kept in mind that these experiments deal with Hans Jenny IJ3 UPTAKE OF Na'BY ROOTS 3- <. y 3 n Uq G re 73 JO tn C Iv» w n n w o n o 3 n> 3 c 13 W cr o o a 6F72/fF £r NH4 BY ROOTS. ME. PER GM o 3 W 3 C o 3 UPTAKE OF K BY ROOTS. ME PER GM ii4 Mineral Nutrition of Plants roots having low-salt status. High-salt roots may release copious amounts of ions to solutions. In view of the fact that a semipermeable membrane inserted between root and clay prevents contact depletion, it appears impossible to resolve this type of root behavior in clay systems in terms of conventional soil solution theories. With Ratner (39, 40) we believe that desorption of root-ions by clays constitutes an important plant physiological process occurring in soils. Uptake of radioactive coliimbium (Cb95) Columbium added to clay behaves as an insoluble compound. At a pH value of 1.0 and at pH values commensurable with root activities, no columbium can be detected in the intermicellar liquid. Yet, as Jacobson and Overstreet (14) have shown, dwarf pea plants success- fully compete with the clay for columbium and accumulate it in the root, and, to a small extent, in the top. These findings have been con- firmed with carrots by Vlamis and Pearson.* We cannot at present comprehend the transfer of columbium on the basis of the carbon dioxide-solution theory. Utilization of nonexchangeable potassium Ramona loam contains 76 p.p.m. of exchangeable potassium as determined by leaching with neutral, normal ammonium acetate. Rye seedlings (Neubauer test) will assimilate, within 18 days, 264 p.p.m. of potassium from the same soil. Evidently these plants extract 188 p.p.m. of nonexchangeable potassium. According to Peech (38), many investigators have postulated the following equilibria between the different forms of soil potassium: Nonexchangeable K ^ Exchangeable K „ K in solution The removal of solution-potassium by the plant releases exchange able potassium into the solution which, in consequence, causes conver- sion of nonexchangeable potassium to the exchangeable form. The *J. Vlamis, and G. A. Pearson, "Absorption of radioactive zirconium and niobium by plant roots from soils and its theoretical significance," Science, 111: 112-113 (1950). Hans ] amy 115 practical validity of the equation in its application to the Neubauer test rests on the rate of conversion of nonexchangeable potassium. Unpublished, exhaustive studies by D. E. Williams show that 30 days of continuous leaching of Ramona soil with various dilute acids (pH 3-6) will liberate only a fraction of the potassium extracted by the Neubauer seedlings in 18 days. Distilled, carbon dioxide-saturated water, passing through 25 grams of soil, freed of exchangeable potas- sium, contains about 0.10 mg. of nonexchangeable potassium per liter, apparently irrespective of the rate of flow. It is highly questionable (/■?) that plants will grow satisfactorily in a solution containing only 0.10 p.p.m. of potassium, even if the solution is continuously renewed. Roots used in the Neubauer test must be endowed with a mechanism of potassium extraction vastly more powerful than carbon dioxide leaching. In the author's opinion, the carbon dioxide theory of liberation of adsorbed ions has generally been overrated. To use a metaphor, we have accepted its application blindly, without asking the plant how far it lives up to its tenets. Besides, the theory is unimaginative; soil water is readily enriched in carbon dioxide and little room is left for indi- viduality of plant behavior. We must look for additional theories of plant nutrition in soils. One of these is the contact exchange theory. THE CONTACT EXCHANGE THEORY The contact exchange theory (/a) describes a mechanism for reac- tions between adsorbents, or solids in general, without the participation of free electrolytes. It was deduced from theoretical considerations concerning the nature of ionic surfaces. Specifically, it rests on the concept of overlapping oscillation spaces of adsorbed ions, or, in another Cloy 1 ~~ Clay Figure 3. Left: Model of ion exchange in salt solution. Right: Model of contact exchange between clay particles. The dashed lines signify overlapping oscillation volumes (75). In these models attention is focused on individual ions. n6 Mineral Nutrition of Plants terminology, on ion redistribution within intermingling electric double layers (Figures 3 and 4). Accordingly, the contact theory embraces all colloidal systems, or all ionic surfaces, not only soil colloids and plant roots. To illustrate the modes of transfer of cations from one negative surface to another let us add radioactive K*-colloid to a sol containing nonradioactive potassium colloid. The two possible types of mechanisms of cation transfer occurring in the mixture may be symbolized as follows: Initial state: Intermedi ate states: J \ Final state: Solution exchange Contact < exchange K* K + HOH + K* K K* K + K*OH + H K K* K + KOH + H K* K* K + HOH + K K* clav K* clay K clay clay K* > K* K clay clay clay clay clay t K clay clay (- clay K K* At equilibrium both particles contain one K*. Analysis of the initial and final states will not tell which mechanism was operative in the transfer of K* from one particle to the other. Of course, if the colloidal particles are prevented from coming close together, only the solution exchange can explain the transfer. Whenever it is impossible to rule out experimentally one of two Hans Jenny ny rival theories, a choice may be made on the basis of the intrinsic value of the theories themselves. There is one important mechanistic aspect which renders the contact theory particularly attractive. Between negative surfaces (negative inner layers), transfer of cations may be accomplished without the aid of anions. Conversely, between Distance Figure 4. Model of contact exchange between inter- mingling ion swarms. Two parallel negative clay plates with undisturbed positive ion swarms on the outer sur- faces and interpenetrating mixed swarms between the plates. positive surfaces (positive inner layers), anions may be transferred without cations. Generally speaking, an ion is transferred without an accompanying partner. Although water is usually present, it is not essential. Theoretically, contact exchange may occur in any medium, provided an electric double layer exists. Sengupta (42) measured ammonium-sodium exchange between 2.02 m.e. ammonium colloid (Ion-X, < 0.1 mm. particle size) and 1.50 m.e. sodium colloid (Amberlite, > 0.5 mm. particle size) in 100 cc. of various media. The transfer of ammonium from Ion-X to Amberlite was 43.1 per cent in water, 45.4 per cent in methyl alcohol, and 21.7 per cent in benzene. In the latter system the rate of transfer was slow and equilibrium probably had not been attained. Ion exchange must have been primarily by contact. In applying the idea of contact exchange to the mineral nutrition of plants in soils, it is postulated that ion swarms of the root and of the soil particles intermingle. Transfer of ions is accomplished without the aid of the soil solution. The uptake of adsorbed ions by contact is n8 Mineral Nutrition of Plants largely independent of the water content of the soil and the transpira- tion stream of the plant. To a certain extent, it will occur at o° C. when metabolic activities are near a standstill. It may even be envisaged that the entire plant is permeated by electric double layers and ion swarms, and that ions may possess locomotion in all plant parts by virtue of contact exchange and surface migration (20). MORPHOLOGIC ASPECTS OF ROOT-SOIL CONTACT EFFECTS The contact theory does not invoke the presence of carbon dioxide. Upon contact, the potassium of the clay and the hydrogen of the root directly exchange positions (Figure 5). While the contact theory dis- Root CONTACT INTAKE CONTACT DEPLETION Figure 5. Schematic representation of contact intake and contact depletion of cations. penses with carbon dioxide it must, on the other hand, assign to the root, properties of an ionic adsorbent, that is, exchange spots. Moreover, the ions assigned to the exchange spots must oscillate sufficiently far to interact with the oscillating ions of the clay particles. It is here that some plant physiologists voice their objections. Wan- ner (47) writes "Dieser Mechanismus . . . ist wenig wahrscheinlich, indem dabei die geringe Kontaktflache zwischen Meristem und Bodenpartikel und die mit Wasser imbibierte Zellulosemembran zwischen den beiden efrektiven Adsorptionsflachen nicht beriicksichtigt werden." The first objection stresses the small contact zone between meristem region and soil particles. Wanner bases his argument on the salt absorp- tion studies of Overstreet and Jacobson (34) which, at o° C, indicate preferential, strong cation absorption in the meristem region which Hans Jenny 119 corresponds to a band having a width of about 1 mm.* Now, one square millimeter of root surface will accommodate 10s clay platelets having an edge length of 100 mu. On each of these clay particles the surface facing the root contains 6000-7000 exchangeable monovalent cations. Accordingly, in a clay soil the ion concentrations surrounding one square millimeter of root surface is tremendous and contact feeding cannot be a limiting factor in supplying ions to the root surface. The second objection of Wanner, the existence of an inert cellulose membrane between the clay surface and the outer surface of the cytoplasm, is also voiced by Lundegardh (29). The thickness of the cellulose wall is quite variable. Lundegardh quotes a range of thickness of 0.1-3.0U or 1000-30,000 A. Frey-Wyssling (10, 11) gives a value of 0.5U for roots. It is not inconceivable that the electric double layers of the cytoplasm and the clay particle are suffici- ently diffuse to penetrate a thin cell membrane from both sides and intermingle within the intermicellar spaces. It is questionable, however, whether this direct contact exchange could bridge cellulose walls which exceed i|j in thickness. Considerable evidence is on hand which refutes the idea of an inert cell membrane. First of all, pure cellulose is not inactive. It has definite, though small, cation exchange properties associated with acidic groups (43). Second, it is very unlikely that the cellulose wall consists of pure cellulose. Frey-Wyssling in his detailed discussion of properties of cell walls reports that the cellulose membrane is permeated by intermicellar spaces which contain pectic substances, lignin, hemicellulose, and mineral substances. Of these, pectic substances are probably most abundant. They have pronounced cation exchange properties. It is also possible that active protoplasmic strands extend into some of the chan- nels of cell membranes of growing root tips. In the light of these considerations, the fine structure of cell walls favors contact exchange rather than disfavors it. The barley roots used by Jenny and Overstreet have a cation exchange capacity (ammonium acetate method) of 11.0 m. e. per 100 grams of dry roots. The precise seat of the exchange spots is not known. How- * According to personal communication by Overstreet, at room temperature nutrient absorption is not restricted to the meristem region. 120 Mineral Nutrition of Plants ever, if living barley roots are immersed in a positively charged iron hydroxide sol, the root surface becomes coated with positive colloids. As this reaction resembles the mutual flocculation process of oppositely charged colloids, the existence of negative charges on the root surface is strongly suggested. Good evidence for the existence of hydrogen ion swarms surrounding the root surface is provided by the suspension effect of roots. Living roots, preferably starved of cations, are washed with carbon dioxide- saturated water (pH 4.1). If the calomel electrode of a glass electrode- pH-meter (e.g., Beckman) is gently but firmly pressed against the root mass immersed in water, the instrument records potential differences which correspond to acidities lying between pH 3-4, and often between PHYSIOLOGIC ASPECTS OF ROOT-SOIL CONTACT EFFECTS The contact theory distinguishes between contact intake and contact depletion of plant nutrients, as illustrated in Figure 5. From clays coated exclusively with potassium ions, excised barley roots sorb large amounts of potassium (Table I). On the other hand, clays devoid of potassium in the ion swarm rob the plant of potassium ions. The de- sorption of root ions bears no simple relationship to pH, and it is not the result of root injury. Needless to say, it cannot be detected if root and clay are separated by a semipermeable membrane (21). There appears to exist for each type of ion on the soil colloid a critical degree of saturation at which the root neither gains nor loses the ion in ques- tion. This null-point would depend on the amount and nature of the clay and on the physiological condition of the plant. As the critical degree of saturation is being approached, the plant suffers severe met- abolic disturbances such as the rosette disease in lettuce (46) which is conditioned by calcium starvation. Ratner (]Q, 40) in two papers published in English reported strik- *D. E. Williams, and N. T. Coleman, "Cation exchange properties of plant root surfaces," Plant and Soil, 2:243-256 (1950). In this connection the follow- ing paper should be consulted: H. Jenny, T. R. Nielsen, N. T. Coleman, and D. E. Williams, "Concerning the measurement of pH, ion activities and mem- brane potentials in colloidal systems," Science, 112:164-167 (1950). Hans jenny 121 ing instances of depletion of ions from barley plants at different stages of growth. Loss of cations to clays containing 30 per cent of calcium and 70 per cent of sodium was not restricted to roots; it extended to stalks, leaves, and even to ears. While loss of nutrients to clays is an observation of long standing (/, 18,23), it nac' never been systematically explained. Now, in the light of the colloid interaction theory it reveals itself as the legitimate partner of contact sorption. TABLE I Gains and Losses of Potassium by Barley Roots in Various Clay Systems Ions on Clays, milliequivalents suspension in 3 liters of Increase or Decrease of K Con- Final pH of Suspen- sion K H Na NH4 Ca Total D* tent of Root; 25 0 0 0 0 25 100% + 44-7% 6.48 5 20 0 0 0 25 20 +28.5 4.40 0 30 0 0 0 3° 0 -66.0 3-52 0 0 9.9 0 0 9.9 0 -13.2 4.70 0 0 0 0 0 0 9.9 0 0 6.8 9.9 6.8 0 0 -32-4 -5-8 5-M 6-43 Hyd -ochloric acid -5-i% 4. 10 D = degree of potassium saturation In presence of clay strongly pronounced counter migration of cations can be demonstrated. From potassium-hydrogen clays (pH 4.4) having not too low degrees of potassium saturation, barley roots accumulate potassium and increase their potassium content by 28.5 per cent. At the same time the roots lose calcium ions, to the extent of 22.2 per cent. From calcium-hydrogen clays (pH 5.25), roots take up calcium ions (6.4 per cent), but experience a loss of potassium ions amounting to 19.4 per cent. With radioactive indicators it is possible to prove that a given ion species simultaneously moves into the root and out of the root. We must think, therefore, in terms of net accumulation and net depletion. 122 Mineral Nutrition of Plants EXAMINATION OF IONIC ENVIRONMENT OF ROOTS AND THE MEASURING OF CONTACT UPTAKE In a soil pore filled with water the liquid phase contains dissolved electrolyte (HoCO:;, Ca(NO.;)2, K2S04, etc.) and, near the surface of the solid phase, the swarm of adsorbed ions. Figure 6 schematically de- picts a root segment — stripped of its electrical properties for the sim- plification of drawing — immersed in a region consisting of a potassium Root segment ^®©^ ® © ® © © © ® ® ® © ® ^®\r ® (ci ® ® ® m k) ®\ ■ \-Root segment Figure 6. Root segment immersed in an ionic environ- ment consisting of ion swarm K (light circles) and KC1 (heavy circles). The ions are meant to he on the outside of the root segment. ion swarm and dissolved potassium chloride. According to the contact theory both forms of potassium are available to the plant, in fact, the root would not be able to distinguish between the two potassium spe- cies. The classic soil solution theory, on the other hand, would not recognize the potassium of the ion swarm as a direct source of potas- sium. As Mattson (_?/) recently put it: the plants feed directly upon XK (K of the free salt) and only indirectly upon ZK (K in combination with clay) ; the latter must be replaced by another cation before it is made available. To experimentally ascertain the reactivity of adsorbed potassium, it is necessary to separate the potassium chloride solution from the ion swarm by centrifugation or ultrafiltration. If root behavior in the origi- Hans Jenny 123 nal, mixed system differs from the root behavior in the extract, a con- tact effect is strongly suggested. The required extraction of the "intermicellar liquid" is often difficult to achieve. Seemingly clear ultrafiltrates, obtained in high-speed cen- trifuges, may reveal numerous fine colloidal particles if viewed in the ultramicroscope. In addition, there are certain aspects connected with Donnan equilibria which must be taken into consideration. The situ- ation is further complicated by carbon dioxide evolution of the roots. Prior to separation the system has to be saturated with carbon dioxide to simulate conditions of root environments. TABLE II Uptake of Radioactive Rubidium by Barley Corresponding Supernatant Roots from Liquids (20) Rb-H-Clays and Degree of Rb Saturation Intake of Rb*, Expressed as Counts per Minute Increase of Clay Sol over Concentration of Sol From clay suspension V^dl uu 11 From bicarbon- Dioxide ate solution Extract 4-94% 2-47 1.65 0.99 i-o% 2.0 3.0 5.0 18.5 24-5 27-5 30 6 12.5 20 30 208% 96 38 0 The soil solution theory contends that any ion removed from the soil solution by plants is replenished by the solid phase. Some investigators even assume that the composition of the soil solution remains constant. Since rate studies also require constant composition of media, the ex- tracted intermicellar liquid must be offered to the roots in such large amounts that the contents remain unchanged during the absorption process. Only two experiments have been published {20, ^5) which satisfy all of these requirements: the sorption of radioactive rubidium as a func- tion of the degree of saturation, and the sorption of radioactive sodium from various clay concentrations. Table II reports the uptake of radio- active rubidium from rubidium-hydrogen clay suspensions and rubid- ium bicarbonate solutions, the latter having the same rubidium con- I24 Mineral Nutrition of Plants centration as the carbon dioxide-saturated supernatant liquid of the rubidium-hydrogen sol. In spite of possible contact depletion by hydro- gen clay, the roots secure more rubidium from the clay sol than from the artificial intermicellar liquid. As roots were in contact with solution and suspension for only one second, ion accumulation inside the root was largely excluded. We are probably dealing here with an exchange involving outer root surfaces only. Figure 7 portrays the sorption of radioactive sodium by groups of 21 intact barley plants during a period of 10 minutes as a function of 120 100- 80 § <0~ 60 k 40 ki 20 3 Na-Clay suspension % 0 Intermicellar liquid C02 saturated 10 20 30 40 50 ME Na-CLAYPER LITER Figure 7. Demonstration of direct utilization of adsorbed ions by roots. sodium clay concentration. The upper curve shows the uptake of sodium from a large volume of sodium clay suspension. The lower curve indi- cates the sodium sorption from a large volume of the corresponding intermicellar liquid, saturated with carbon dioxide prior to separation. During the process of ion accumulation by roots, clay sols, as well as sodium bicarbonate solutions, were continually aerated. In our opinion, the difference in the two curves demonstrates direct utilization by roots of adsorbed sodium. Hans Jenny 125 EQUILIBRIUM STUDIES ON ION EXCHANGE BETWEEN ADSORBENTS Symbolically, we may formulate the relationship of contact intake and contact depletion between plant roots (R) and clay particles: Vh + K clay ^ V K + H clay | The question whether the reaction will tend to the left (contact deple- tion) or to the right (contact uptake) depends on the amounts of the two adsorbents and on the forces with which the various ions are held to the two surfaces. In approaching a quantitative elucidation of the above process, it appears advisable to first study in detail the exchange reaction between two adsorbents in vitro. Let us consider the exchange of the ammonium and sodium ions, between two different adsorbents, either by solution or by contact, as represented by the following system: Ion-X NH4 + Na Amberlite Ion-X Na + NH4 Amberlite Amberlite and Ion-X are artificial cation exchange resins. It is pos- sible to study the reaction quantitatively by mixing coarse Amberlite particles (> 0.5 mm.) with fine Ion-X particles (< 0.1 mm.). After a shaking period of several days the two adsorbents are separated by wet sieving and analyzed individually. In the writer's laboratory, Dr. K. Sengupta has shown that the above reaction gives a characteristic exchange constant according to the equation: Na Ion-X • NH4 Amberlite .46 (experimental) NH4 Ion-X Na Amberlite It is possible to calculate this exchange constant without actually mix- ing the adsorbents, provided the ordinary base exchange reactions be- tween the adsorbents and neutral salt solutions are known. Consider the two connecting vessels of Figure 8. Initially the stopcock is closed. The left hand vessel (system X) contains coarse ammonium- Ion-X and sodium chloride. At equilibrium we have: 126 Mineral Nutrition of Plants Ion-X NH4 + NaCl ^ Ion-X Na + NH4C1, or Na Ion-X (NH4) = ^x = 0.793 (experimental) NH4 Ion-X (Na)x 1 1 J L NaCl NH4CI Ion XI NH4 Na NaCl NH4CI ^-—1 NH4 ^ Na System X System A Figure 8. Device for deriving exchange equa- tions between adsorbents. This equation holds for wide variations in the amounts of sodium chloride or adsorbent used. The right-hand vessel (system A) contains coarse ammonium-Am- berlite and sodium chloride. The equilibrium equation assumes the form: Na Amberlite (NH4) = /(A = 0.532 (experimental) NH4 Amberlite (Na) Again variations in the amounts of reacting substances do not sub- stantially alter kA. Now we open the stopcock and let a new equilibrium be established. Whatever the initial salt concentrations were, we have now: (NH4C1)X = (NH4C1)A and (NaCl)x = (NaCl)A By mere algebraic rearrangement we obtain: Na I Ion-X | NH4 | Amberlite ^x 0.793 K = 1.49, Q.E.D. NH4 I Ion-X Na Amberlite ^A 0.532 Hans Jenny 127 In terms of the oscillation volume theory (75), the constant K com- prises four oscillation volumes, conditioned by the two different ions each being held on two different adsorbents. For the system under con- sideration: K = l_ v N V* JfNa _ ^NH4_ For interacting mono-divalent ions we write: NH4 Ion-X NH, -f- Ca Amberlite Ion-X Ca + NH4 NH4 Amberlite Applying the base exchange equations of Krishnamoorthy, et al. (26), the following equilibrium equation may be derived: Ca(ii Ca + NH4) I Ion-X (nh4 Amberlite NH4 Ion-X Ca(i| Ca + NH4) Amberlite -= k The symbol kA represents the equilibrium constant for the system Ca-Amberlite -f- 2 NH4C1 and kx the constant for Ca-Ion-X -j- 2 NH4C1. In numerous experiments Sengupta has shown that the exchange of ions of equal valency (NH4-Na, NH4-K, Ba-Ca) between two ad- sorbents is normal, that is, the observed equilibrium constant is identi- cal, within experimental error, with the calculated constant (Table III). The ammonium-cesium pair possibly constitutes an exception. This identity does not exist for ion pairs of unequal valency (Na-Ca, NH4-Ca). If large particles of ammonium-Amberlite are mixed with small particles of calcium-Amberlite (same type of adsorbent), the ex- change constant should be 1, but according to Table III it is over twice as large — namely, 2.1 1. In every case listed in Table III, the observed exchange constant differs profoundly from the calculated one. Interestingly enough, the abnormality disappears when free salts are added to the adsorbent mixtures. Now, the observed exchange constant agrees with the calculated one. At present we do not know whether this peculiar behavior of the mono-divalent ions is the result of experi- mental technique or of colloid interaction. 128 Mineral Nutrition of Plants TABLE III Exchange Constants Between Adsorbents (K. Sengupta) Adsorbent I Adsorbent I K K Ob- tained Calcu- lated Particle Particle Type size Type size NH4-Ion-X >o.5 mm. Na-Ion-X 0-5 mm. Ba-Ion-X o.5 mm. NH4-Ion-X o. 5 mm. NH4-bentonite <0.I> 0.05 mm. 0.40 o-39 NH4-Ion-X >o. 5 mm. Cs-bentonite < 1 ju OI34 0. 105 NH4-Amberlite >o.5 mm. Ca-Amberlite o. 5 mm. Ca-Amberlite o. 5 mm. 0.415 1 .00 NH4-Amberlite >o. 5 mm. Ca-bentonite < 1 JU 18.70 6.92 NH4-Ion-X 0 . 5 mm. 1. 91 3.00 TIME RATES OF ION TRANSFER BETWEEN ADSORBENTS The aforementioned adsorbent-adsorbent studies deal with equilib- rium states. In so far as alkali metal ions are concerned, the transfer may be visualized as taking place through solution or by contact. In either case, the amount of net transfer is the same. It is, however, highly probable that the rate of transfer is greatly accelerated by contact ex- change. In fact, conditions exist under which solution transfer is so slow that for practical purposes it is nonexistent. Good illustrations are provided by the behavior of divalent iron adsorbed on bentonite (20). More recent experiments furnish information on columbium and non- exchangeable potassium. Experiments with columbium (K. Sengupta) Radioactive columbium, Cb95, carrier-free and dissolved in 3N sulfuric acid, was added to Amberlite and to clay suspensions. After excess acid had been washed out the supernatant liquid was completely free of columbium. A 0.17 per cent suspension of columbium-hydrogen clay had a pH of 3.0. Adding hydrochloric acid to bring the suspension to 0.025 AT hydrochloric acid did not release any columbium; neither Hans Jenny 129 did sodium chloride, calcium chloride (0.08 N), or calcium hydroxide. According to Klimenko and Syrokomskii (24), columbium precipitates above pH 0.4. Whereas dilute acids and neutral salts are unable to release colum- bium, contact will do so. The transfer is reversible; columbium will move from fine clay to coarse Amberlite, and vice versa. During four days of shaking of adsorbent mixtures the following amounts of colum- bium were transferred from left to right: Cb-H-bentonite + H-Amberlite 8.6 per cent Cb-Na-bentonite -+- Na-Amberlite 5.8 per cent Cb-Ca-bentonite -f- Ca-Amberlite 1.7 per cent Cb-Na-Amberlite + Na-bentonite 6.0 per cent Cb-Rb-Amberlite -f- Rb-bentonite 10.4 per cent Time curves indicate that in these experiments equilibrium had not yet been reached. Although the rate of transfer is slow, it is in the realm of practical significance. Release of nonexchangeable potassium D. E. Williams prepared mixtures of Ramona soil and Ion-X ad- sorbent. After one week of shaking the mixtures were leached with ammonium acetate and potassium was determined in the leachate. According to Table IV much larger amounts of nonexchangeable po- tassium are liberated by contact with adsorbents than by leaching with the corresponding neutral salts. The liberation of fixed potassium be- comes especially noticeable when hydrogen-Ion-X or hydrogen-Amber- lites are used. Conceivably the hydrogen ion swarm surrounding the root likewise acts as an efficient releaser of nonexchangeable potassium. It appears plausible that the contact effects involving columbium and nonexchangeable potassium explain, partially at least, the root behavior discussed on pages 114 and 115. EXTENT OF CONTACT EXCHANGE IN SOILS In any soil both solution and contact mechanisms will be operating. As far as macronutrient cations are concerned, the soil solution mecha- nism would be expected to predominate in sandy soils, whereas in clay 130 Mineral Nutrition of Plants soils contact would be the decisive factor. Also, the lower the salt con- tent of the soil solution the greater will be the contribution of contact exchange. For those micronutrient cations, including iron, which are largely insoluble at higher pH values, contact exchange may well be the dominant mode of acquisition by roots. TABLE IV Extraction of Nonexchanseable Potassium from Ramona Soil Nonexchangeable Potassium Extractant and Method Extracted, p.p.m. Neubauer rye seedlings test, 18 days 188 Leaching with N NH4-acetate, 8.5 liters in 35 days . . o Leaching with C02 saturated water (pH 4.0) 8.2 liters in 35 days !5 10. o liters in 14 days 27 Leaching with N/ 10 NaCl, 1.25 liters in 7 days . ... 23 Shaking 5 g. of soil for 7 days with 25 m.e. NH4-Ion-X, pH = 7.6 . . 68-114 Shaking 5 g. of soil for 7 days with 25 m.e. Na-Ion-X, pH = 7.8 1 10-152 CONCLUSION The contact theory of mineral nutrition of plants in soils is only a theory. If the scientist-philosopher Campbell (6) is right, we shall never know whether or not it portrays the truth, for as he says, theories can never be proven, they can only be disproven. However, the contact theory is a productive and stimulating theory. It assigns to the root surface the role of active, individualistic, genetic, and physiologically conditioned participation in the liberation of ad- sorbed nutrient ions. It enables the plant to feed directly upon the solid phase. It insists upon the interplay of soil colloids and plant colloids.* In short, it discloses new vistas on the behavior of plants in soils. *As Dr. Norman points out elsewhere in this volume, contact exchange should likewise include the surfaces of the microbial soil population. Hans Jenny 131 REFERENCES 1. Albrecht, W. A., and McCalla, T. M., Am. /. Botany, 25:403 (1938). 2. Allaway, W. H., Soil Sci., 59:207 (1945). 3. Arnon, D. I., and Grossenbacher, K., Soil Sci., 63:159 (1947). 4. Ayers, A. D., Thesis, Univ. Calif. Berkeley (1939). 5. Cameron, F. K., The Soil Solution (Easton, Pa., Chem. Publ. Co., 1911). 6. Campbell, N. R., Physics, The Elements (Cambridge Univ. Press, 1920). 7. Comber, N. M., /. Agr. Sci., 12:363 (1922). 8. Elgabaly, M. M., Jenny, H., and Overstreet, R., Soil Sci., 55:257 (I943)- 9- , and Wicklander, L., Soil Sci., 67:419 (1949). 10. Frey-Wyssling, A., Die Stoffausscheidung der hoheren Pflanzen (Berlin, }. Springer, 1935). ir. , Erniihrung und Stofficechsel der Pflanzen, (Zurich, Biicher- gilde Gutenberg, 1945). 12. Hoagland, D. R., and Broyer, T. C, Plant Physiol., 11:47 (J93^)- 13. , and Martin, J. C, Proc. First Intern. Congr. Soil Sci., 3:381 (1928). 14. Jacobson, L., and Overstreet, R., Soil Sci., 65:129 (1948). 15. Jenny, H., /. Phys. Chem., 40:501 (1936). 16. , and Ayers, A. D., Soil Sci., 48:443 (1939). 17. Jenny, H., and Cowan, E. W., Science, 77:394 (1933). 18. , Z. Planzenerniihr. Diingung u. Bodenl^., A 31:57 (1933). 19. Jenny, H., and Overstreet, R., Proc. Nat. Acad. Sci. US., 24:384 (1938). 20. , /. Phys. Chem., 43:1185 (1939). 21. , Soil Sci., 47:257 (1939). 22. Jenny, H., Overstreet, R., and Ayers, A. D., Soil Sci., 48:9 (1939). 23. Kelley, W. P., Proc. First Intern. Congr. Soil Sci., 4:483 (1928). 24. Klimenko, N. G, and Syrokomskii, V. S., Chem. Abstracts, 43:4083 (1949). 25. Kossovitch, P. S. (1902); see Parker (57). 26. Krishnamoorthy, C, Davis, L. E., and Overstreet, R., Science, 108: 439 (1948). 27. Liebig, J. v., Ann. Chemie u. Pharm., 105:109 (1858). 28. Loosjes, R., Landbouw\und . Tijdschr., 52:836 (1940). 29. Lundegardh, H., Soil Sci., ^-\:iyy (1942). 30. Marshall, C. E., Soil Sci. Soc. Am. Proc, 7:182 (1942). 132 Mineral Nutrition of Plants 31. Mattson, S., Ann. Agr. Coll. Sweden, 12:119 (1944). 32. Mehlich, A., and Colwell, W. E., Soil Sei. Soc. Am. Proc, 8:179 (1944). 33. Overstreet, R., Broyer, T. C, Isaacs, T. L., and Delwiche, C. C, Am. f. Botany, 29:227 (1942). 34. Overstreet, R., and Jacobson, L., Am. f. Botany, 33:107 (1946). 35. Overstreet, R., and Jenny, H., Soil Sci. Soc. Am. Proc, 4:125 (1939). 36. Overstreet, R., Ruben, S., and Broyer, T. C, Proc. Nat. Acad. Set. U.S., 26:688 (1940). 37. Parker, F. W., Soil Sci., 24:129 (1927). 38. Peech, M., Diagnostic Techniques for Soils and Crops (Washington, D. O, Am. Potash Inst., 1948). 39. Ratner, E. I., Compt. rend. acad. sci. U.R.S.S., 42:313 (1944). 40. , Compt. rend. acad. sci. U.R.S.S., 43:126 (1944). 41. Schachtschabel, P., Erniihr. Pflanze, 35:132 (1939). 42. Sengupta, K., Thesis, Univ. of Calif. Berkeley (1949). 43. Sookne, A. M., and Harris, M., in Ott, E., ed., Cellulose and Cellulose Derivatives (New York, Interscience Publ. Co., 1946) 113. 44. Truog, E., Proc. First Intern. Congr. Soil Sci., 3:628 (1928). 45. Vageler, P. W. E., Kationen- und Wasserhaushalt des Mineralbodens (Berlin, J. Springer, 1932). 46. Vlamis, J., Soil Sci., 67:453 (1949). 47. Wanner, H., Vierteljahrsschr. naturjorsch. Ges. Zurich, 43:99 (1948). CHAPTER 6 The Effect of Soil Phys- ical Properties on Nutrient Availability J. 6. PAGE and G. B. BODMAN S. ince the beginnings of plant science the major emphasis has been placed upon the chemical properties of the soil and plant nu- trients. Liebig stated in 1840: "The crops on a field diminish or increase in exact proportion to the diminution or increase of the mineral sub- stances conveyed to it in manures." This view has been widely accepted, and can easily be understood in light of the very favorable yield in- creases usually obtained when fertilizers are applied to the soil. The proper use of fertilizers has become an accepted and necessary part of farm operations. There are, however, other factors of importance in crop production. This has long been recognized but comparatively little attention has been paid to the physical properties of the soil, particularly as they affect plant growth and fertilizer response. Studies in soil fertility and soil physics have proceeded concurrently with but little appreciation of the significance of the interaction of soil physical properties upon the better understood and more intensely studied chemical properties of the soil. Many accept the physical properties of a soil as being extremely important but apparently feel that little or nothing can be done to effect changes or improvements in such properties. With some soils, particu- larly the well-drained sandy soils, such a view may be partly justified since they have a textural porosity which permits favorable water and air movement. For a great many very important agricultural soils, how- ever, the physical behavior is dependent upon structure and aggregation and these not only play a very important role in crop production but 134 Mineral Nutrition of Plants are subject to rapid change with time or treatment. There appears to be little appreciation of this fact. This is indicated in part by the very common tendency to depend upon fertilizers alone with little thought given to physical factors when it is desired to increase the productive capacity of a soil. In recent years there has been an increasing appreciation on the part of some workers of the importance of the physical properties of the soil in crop production. The reasons for this are probably threefold: (a) with general adoption of high yielding varieties of crops and heavy fertilization rates, frequent failures to obtain the expected yield in- creases have shown that factors other than fertility often limit yields; (b) with continued cultivation many of our soils have lost or are losing much of their organic matter with a corresponding deterioration of favorable soil structure, so that physical properties are becoming poor enough to seriously limit plant growth, even though fertilizers may be liberally supplied; and (c) under semiarid and arid conditions, where initial content of soil organic content was low, a change from native vegetation growing on undisturbed soil under low rainfall to intro- duced field and orchard crops growing on frequently tilled, irrigated soil, has been observed to be associated with distinct structural changes. Recent well-publicized attempts to make record-breaking corn yields have shown very strikingly that for high yields to be obtained, the physical, as well as the chemical factors must be at their optimum. It has also become quite apparent that unfavorable physical properties have profound effects upon the chemistry of the soil and upon availa- bility of plant nutrients, both native and added. Little direct work has been done on this subject, so that there is need for a general appreciation of this lack in our knowledge of the soil-plant system, and for a con- certed attack on the problems by both soil and plant scientists. It is the purpose of this paper to point out the importance of the physical properties in crop production, the ways in which physical properties change or affect nutrient availability, and the need for a general reali- zation of the inadequacy of the strictly chemical approach and for the study of the soil as a system. Page and Bod man 135 SOME FACTORS ESSENTIAL FOR PLANT GROWTH The soil factors which are essential for normal plant growth are: (a) a favorable soil reaction and an adequate supply of all the essential nutrients, (b) a favorable supply of water, (c) adequate oxygen, (d) favorable temperature, and (e) friability or looseness of the soil so that roots are not restricted in their free growth and development. Of these factors it is usually assumed that the supply of nutrients and the reac- tion can be controlled through addition of fertilizers and lime, but where the other factors are unfavorable, added nutrients may not be effective and hence, in the broad sense of the term, can be considered as being unavailable. The role of the water supply in affecting availa- bility of nutrients will be the topic of another paper in this symposium and will not be discussed here. In discussing the remaining topics, "availability" will be treated in the sense indicated above, i.e., whenever plants fail to assimilate and utilize nutrients, the nutrients will be considered as being unavailable even though they may be in solution or appear to be readily available from chemical tests. It is realized that this is an unorthodox usage of the term "availability" but, as will be seen later, the effect of physical properties of the soil on plant growth and utilization of nutrients may be very significant. In the opinion of the authors too much reliance on a strictly chemical definition of availability of plant nutrients is mislead- ing and only through adoption of the broader definition and all that it implies can we gain a proper understanding and interpretation of plant growth and fertilizer response. When soil physical properties are unfavorable, nutrient availability may be reduced in any or all of the following ways: by restriction of root extension, thus limiting the volume of the soil which the roots can penetrate and the number of particles with which root contact can be made; by affecting the chemistry of the soil so that nutrients are no longer in soluble form; by limiting the metabolic activity of roots so that they can no longer function properly in assimilating the soluble or exchangeable nutrients in the soil. 136 Mineral Nutrition of Plants SOIL PHYSICAL PROPERTIES The physical properties of a soil play a significant role in plant growth by controlling the air, water, and, to a certain extent, nutrient supply to the roots of growing plants. In considering these properties we are primarily concerned with the pore spaces between the soil particles because it is here that roots exist, and the necessary water and air move- ments occur. The proportion of the total soil volume made up of pore space, the size, and shape of the pores are all extremely important. These characteristics in turn are governed by two properties: the par- ticle-size distribution and the arrangement of the soil particles, particu- larly as they may be clustered into larger aggregates or crumbs. Particle-size distribution and dependent properties That quality of a given sample of soil taken from the soil profile, which may be expressed as a particle-size distribution curve, probably represents the closest approach to an intrinsic physical soil property. But it must be understood that in order to obtain reproducible results particle-size measurements must be made according to standardized arbitrary methods of soil pretreatment and dispersion. For purposes of obtaining expressions of "ultimate" particle-size dis- tribution, dispersion methods commonly in use aim at the highest pos- sible dispersion short of actual solution. Expressions having this basis are now used in the choice of soil texture names. However, C. L. Clark (29) pointed out in 1933 that at that time there was no proof in support of the correctness of the assumption that a unique particle-size distri- bution exists for a given soil sample and, to the authors' knowledge, none has been presented in the subsequent fifteen years. From the published experiences of investigators in soil physics it must be concluded that most, if not all, other physical properties that have been examined are remarkably susceptible to modification by external factors, and to external forces of one kind or another which have been brought to bear upon the soil. The physical factors which affect plant-soil interrelations, therefore, are— with the qualified excep- tion of particle size of the solid phase— notably subject to quantitative change, brought about both by natural events and by the practice of soil management. Under field conditions, alterations associated with the Page and Bod man 137 passage of time appear to proceed least rapidly with the property of particle-size distribution. It is doubtful that any physical property of soil remains uninfluenced in some respect by the proportions which the soil contains of particles of different sizes, since the "ultimate dispersion" obtainable for a given soil (by the arbitrary methods already mentioned) is itself a manifesta- tion of the particle-size distribution, and the extent to which this limit- ing degree of dispersion is approached is of great significance. Cultiva- tion and management practices, the growing plant, microorganisms and the weather, all tend to shift the degree of dispersion of the soil as displayed at any one instant closer to, or further away from, the state of ultimate dispersion. The physical properties to be discussed may all be shown, under certain conditions, to be functions of the degree of dispersion of the solid phase. Yet, at present, nothing less than a complete distribution curve and a fundamental expression of aggregation status, together with full information concerning its chemistry, mineralogy, micro- biology, and organic matter, will even approximately specify the physi- cal and physico-chemical behavior of the soil under a given sequence of treatments and changes in water content. The functional relation- ships are extremely complex. Role of aggregation The tendency of soil particles to form clusters or aggregates which resist dispersive treatment has received much attention by soil physicists. Measurements of approximate aggregate size distribution by sieving in air, and sieving, shaking, settling, or elutriating in water, have been used to provide numerical expressions of soil structure. Although the results are affected by the method of determination used and no stand- ard method of measurement has yet been adopted, aggregate size dis- tributions have been shown to affect other physical soil properties and also plant growth. Clarke and Marshall (jo, j/) and others have shown the existence of a significant positive effect of clay on water-stable aggregation. They found that a significant decrease in water-stable aggregation* resulted * Aggregation was expressed in terms of the amounts of primary particles smaller than a given size which were present in aggregates larger than that size. 138 Mineral Nutrition of Plants from increasing periods of cultivation of South Australian grassland in the red-brown earth zone. The major part of the decline took place during the first five years out of a total of twenty years of cultivation. Similar observations have been made in this country. The effect of aggregate size upon yield of Synapis alba (white mus- tard) grown in Mitscherlich pots filled with soil aggregates of different size classes was studied by Schuylenborgh (62). The aggregates and crumbs were obtained by screening a cloddy clay soil. With respect to structural state and aggregate size class, Schuylenborgh found that dry weight yields increased in the order: (compact structure) < (4-5 mm.) < (2-4 mm.) < (finer than 1 mm.) < (1-2 mm.) < (crumb structure).* These results are in general agreement with those obtained by Kwasnikov (see Krause, 45) who conducted a somewhat similar experiment. Doyarenko (see 45) and Schuylenborgh (62), respectively, regard the aggregate size classes 1-2 mm. and finer than 1 mm. as highly desirable. Unfortunately, the last category fails to specify the amount of aggregation within this size class. It cannot be said that any one aggregate size class is best for all soils which possess the ten- dency to form stable aggregates. It seems more likely for soils of dif- ferent intrinsic physical and chemical properties that different states of aggregation will be required for the growing plant. In this respect, moreover, different plants may have different requirements. The crumb structure is less permanent than that associated with the water-stable aggregates of size classes < 1 mm. and 1-2 mm. For many soils in certain climates it may be too costly to attempt to create this structural state. Superior yields on the crumb structure and finer aggregates were attributed by Schuylenborgh to the greater surface areas there exposed for root exploration. This may be the explana- tion for the results obtained with this particular soil but, omitting the crumb structure, there is no evidence to indicate that some other size classes than those finer than 2 mm. might not be superior for other soils, e.g., those which form soft aggregates. The influences of particle size upon the availability to plants of soil * Crumbs have weak cohesion and lack distinct dimensions and shape; they possess a honeycomb structure formed by the combination of primary soil parti- cles with larger aggregates (Schuylenborgh). Page and Bod man 139 water may be demonstrated, for soils of different texture, by direct measurement of the wilting point (72) or from moisture potential- moisture content curves, which may be obtained by means of pressure membrane apparatus (5S) and by other methods (72). The permanent wilting percentage and moisture equivalent (a func- tion of particle-size distribution) values published for a number of dif- ferent surface soils by Richards and Weaver ($7) and by Veihmeyer and Hendrickson (69) indicate a high positive correlation. This may be explained by the existence, within soils of high clay content, of very fine pores that retain water against plant withdrawal forces and that are present in greater numbers in the clay-rich soils than in those of lower clay content. Although the functional relationship between permanent wilting percentage and soil is far from a simple one and may involve mineralogical and physical characteristics, the proportion of very fine particles present is possibly the most important single factor affecting this property in nonsaline soils. For a given total water con- tent below the wilting point the energy of water retention generally increases considerably with increasing fineness of texture. Many fine-textured highly water-retentive soils also show a marked tendency to be very plastic and to remain wet for a comparatively long time. Such soils are too frequently worked while they are still wet, with the result that puddling occurs and the soils thus become even less satisfactory for plant growth. Strikingly harmful effects on plant growth have been observed both in the field and the laboratory as the result of soil puddling: the process of mechanically working a soil and altering its structure so that, by dispersion of the aggregated particles, the coarser pores are destroyed. For a definite expenditure of mechanical work of compression and shear, the amount of dispersion of aggregates by puddling treatment and the resultant diminution in apparent specific volume appear to depend upon the moisture status of the soil, the amount of mechanical work already expended in reducing the apparent specific volume, and the particle-size distribution (/j). This last, of course, within the limitations stated earlier, represents the ultimate in mechanical disper- sion for a given soil. It has been shown by Thomas (67), Day (32), and Buehrer and 140 Mineral Nutrition of Plants Rose (19) that, at a given moisture content and between certain mois- ture potential limits, dispersion by puddling diminishes the moisture potential, i.e., moisture retention forces are increased. Although the diminution appears significant, the magnitude of the effect in some cases may be slight (12). McGeorge and Breazeale (5/) expressed the opinion that puddling increased the moisture content at which plants first displayed symptoms of moisture unavailability. Total porosity diminution by puddling proceeds at the expense of the coarser pores, which are destroyed relatively easily. It is largely in these coarser pores that air and water movement occur and that roots find their most favorable environment. The benefit derived from aggregation is largely associated with the increase in the important larger pore spaces in the soil. Sands, of course, may have sufficiently large pores for normal air and water movement but the pore spaces in most silts and clays, if these soils are not aggregated, are usually so small that capillary forces cause them to remain full of water to the exclusion of air. The persistence of smaller pores, despite puddling treatment, can be explained in three ways, according to Rubin (59) : (a) destroyed small pores of a given radius are replaced by others re- sulting from decrease in radius of larger pores; (b) smallest pores are filled with water which resists removal and prevents pore collapse; and (c) small pores owe their existence, in large part, to a more stable particle arrangement in their immediate vicinity, i.e., they may (and probably do) exist in greater abundance within the real or potential aggregates which are present in the soil mass. Increase in relative abundance of very fine pores conceivably may affect water availability to plants, directly, by increasing the energy of water retention and, indirectly: (a) by offering greater resistance to root penetration, (b) by decreasing the supply of necessary oxygen to the roots with the result that roots no longer function normally in absorbing water, and (c) by increasing, over short distances, the re- sistance to water movement toward the plant roots. The mechanisms involved should be given more attention than they have heretofore re- ceived, but the net effect to the plant would be qualitatively similar, namely, the availability of the existing water supply would be dimin- ished. Page and Bod man 141 Penetrability by roots The impenetrability of unfractured, dense rock to plant roots needs no comment. In shallow residual soils plant growth may be seriously restricted by the presence of continuous rock masses. Dense layers of soil also may be expected to resist root penetration provided that the density is sufficiently high or the pores sufficiently small. For example, restricted root growth may be observed over the surface of puddled cultivation-soles in citrus orchards and within crevices over the sur- faces of dense structural columns in the subsoils of some alfalfa fields. Uptake of water and other nutrients will thus be restricted. In Cali- fornia there is now being followed a practice of noncultivation of orchards whereby danger of puddling is diminished and previously puddled soils are allowed, by natural processes of wetting and drying in the undisturbed state, to recover their former more open structure. Root development is thus enhanced. Soils of low total porosity and relatively high microporosity may be almost impermeable to water. The B horizon of the San Joaquin series (California) has been shown (//) in some profiles to have a total porosity (0.26 to 0.34) inadequate to accommodate its moisture equiva- lent water, the magnitude of the latter being determined in the stand- ard way on disturbed and crushed soil lumps. This horizon, assuming its penetrability to both roots and water, would be deficient in the total water available for plant growth if for no other reason than that the upper limit of growth water would be significantly less than in the case of the same material in the less compact state. Field observation reveals very few roots in the B horizon of these soils. An apparent anomaly was observed by Veihmeyer and Hendrickson (yo) in the seasonal change in moisture content of the subsoil of the Bale gravelly loam (California) under grapevines, and in that of a certain primary soil supporting a chaparral plant association. Despite water additions and plant transpiration these subsoils showed but slight changes with time in the content of total water, which suggested lack of absorption by roots. A density of 1.83 g. per cc. was observed in the Bale subsoil. Later work (7/) was conducted in which the penetrability to sunflower roots was examined for several different manually com- 142 Mineral Nutrition of Plants pacted soils. The soils were compacted while moist to densities ranging from 1.46 to 2.06 g. per cc. At the beginning of the experiment they con- tained amounts of water calculated in all but one case to be in excess of their permanent wilting percentages, and in all cases to be well below saturation at the final, increased density sought; but no roots penetrated any of these soils and there was negligible loss of water from them. These authors have also observed little or no water extraction by pine trees, grapevines, fig trees, and chaparral shrubs from subsoils having the densities stated. No one density represented the limiting density for all soils examined, but results suggested that somewhat lower densities prevented root penetration more in fine than in coarse-textured soils. SOIL AERATION Soil aeration is the interchange of the soil atmosphere with the free atmosphere above the soil. Continued soil aeration is essential to re- move the carbon dioxide produced by plants and microorganisms and to supply the oxygen needed by plant roots and the microorganisms in the soil. Obviously this interchange must occur in pores large enough that capillary forces do not keep them filled with water, and which have access to the atmosphere through a network of interconnected, unobstructed pore space channels. It has long been recognized that good soil aeration is needed for normal plant growth, and most farmers or gardeners are of the opinion that one of the chief purposes of tillage or drainage is to assure good soil aeration. In spite of this, comparatively little work has been done on the actual role of soil aeration in plant growth and little is known about the mechanics of the interchange of gases in the soil. Two facts have, however, been quite definitely estab- lished: oxygen is essential for normal root growth and extension, and oxygen is essential to the root if it is to carry out its normal func- tion of absorbing nutrients and water. Much of the work which has been done on the role of oxygen in root growth has been done in solution cultures. Admittedly, solution cul- tures and soil offer quite different environments for the growth of roots. Roots of different plants may develop and function differently in solu- tion cultures than they do in soils. If, however, it is demonstrated that roots require oxygen for their normal functioning in solution culture, Page and Bod man 143 it seems legitimate to conclude that a similar requirement might be demonstrated in soil. Actually there is strong evidence for the essen- tiality of oxygen for root development in either type of environment. Effect of soil aeration on root growth Much of the solution culture work has been done with tomatoes. Arnon and Hoagland (j) obtained an average weight of 12.4 grams per plant for roots in unaerated culture solutions as compared with 19.9 grams per plant for aerated solutions. The yield of tomatoes for the aerated solution was 1.34 times that in the nonaerated solutions. Arrington and Shive (4) obtained from culture solutions dry-weight yields that were increased 210 per cent as the result of aeration. Erick- son (34) too found that aerated plants were almost twice as large as those which were not aerated. Clark and Shive (28) and Stiles and Jorgensen (66) found that the dry weights of aerated roots were 1.6 times the weights of nonaerated roots, but Allison and Shive (2) found this ratio to be only 1.12 for soybeans. Most workers have pointed out that individual crops difTer in their oxygen requirements, but none pre- sent clear-cut evidence that the roots of any plant can grow and func- tion in the complete absence of oxygen. Gilbert and Shive (^5) give data for response to oxygen at various levels for different plants, and Erickson (34) gives an excellent recent review of the effects of oxygen in culture solutions. He found that root growth increased in proportion to the oxygen content of the culture medium. Several observations of rooting habits in aerated versus nonaerated solutions are significant. Shive and his co-workers found that aerated roots were long, slender, and much branched. Erickson observed root- ing habits similar to those reported by Bryant (18) who found that the length of aerated roots in his experiments averaged 37.4 cm., while those of nonaerated plants averaged only 10.9 cm. in length. The non- aerated roots were 15 per cent thicker than those aerated, and it was observed that aerated roots had a longer portion of the root tip over which absorption of water and salts could occur. Similar results have been reported for plants grown in soils. Loehwing (48) found that soil aeration helped to produce plants having larger 144 Mineral Nutrition of Plants tops and roots, and that roots in aerated soils were distinctly more fibrous in character, longer, and more numerous than those from poorly aerated soils. He also found that roots grown in aerated soil had approx- imately twice the surface area of roots on plants grown in nonaerated soil (47). Knight (44) found that artificial aeration of soils increased top and root growth by 30 per cent in pot soil cultures. Baicourt and Allen (5) grew roses in aerated and nonaerated soils. They found that the aerated plants grew 68.4 inches, while the nonaerated plants grew only 37.5 inches in three months. Bushnell (20) aerated the soil in which potatoes were growing by placing a line of perforated tile in the soil. He obtained a yield increase of 15 to 29 per cent for aeration and observed that roots were more abundant around the tile line than in the main body of soil. A somewhat similar result was observed by one of the present authors for heavily watered tomatoes in a com- mercial greenhouse. Here growing roots were found only along the inside surface of the sterilizing tile, or spread out over the surface be- tween the mulch and the soil. Presumably, deficient soil aeration caused by excessive watering prevented normal growth and development of roots in the main body of the soil. Several experiments conducted in Ohio established a significant re- lationship between degree of soil aggregation, pore size, and crop yield, particularly on heavy clay soils. A high degree of soil aggregation or a fairly large proportion of larger pores was interpreted as indicating conditions favorable for good soil aeration. Baver and Farnsworth (6) found that, where total air capacity of the soil was only 3 per cent by volume, sugar beets suffered a 50 per cent loss in stand from black root rot disease and yielded only 2-4 tons. Very much higher yields of sugar beets were obtained after the introduction of cropping and manage- ment practices that increased the proportion of coarse pores present in the soil. Where air space porosity was raised, the loss in stand was re- duced to only 10 per cent and the yield was raised to 12 tons. Page and Willard (5^) found a fair degree of correlation between air space poros- ity and corn yield. Rotations that led to increased soil particle aggrega- tion and, hence, better air space porosity and aeration were associated with large increases in corn yields, with average yields measuring from 49 bushels in 1936 to 84 bushels in 1945. They concluded that crop yields Page and Bod man 14c in their experiments were definitely limited by soil structure and aera- tion. Trogdon (68) found a highly significant correlation between yield of corn and air pore space in the soil, irrespective of the amounts or forms of fertilizer applied, provided essential elements were not lacking. Cannon and his co-workers have conducted outstanding research on root-aeration relationships. Cannon and Free (26) stated: "Increased air supply to roots if not excessive favors root branching and probably accelerates the rate of root growth." They further suggested that aera- tion may be of equal importance ecologically with temperature and water supply. The effect of low oxygen levels and temperature on the growth rate of corn is brought out in Table I which gives some of Can- non's (24) results. TABLE I Effect of Oxygen Levels and Temperature on Rate of Growth of Corn (Cannon 2^) Rate of Growth of Corn With the Following Oxygen Levels in Soil Air Temp. 3 per cent 3.6 per cent 10 per cent normal 1 8° C. 1/3 normal 2/3 normal 200 1/5 normal 3°° i/J6 normal 1/3 normal 9/10 normal It can be seen from Table I that there is a significant influence of temperature upon the relation of roots to oxygen. The effect of low oxygen levels is most pronounced at higher temperatures so that for best growth rates during hot weather, soil aeration must be excellent. Cannon (22,25) a^so studied several species of plants differing in natural habitat and concluded: (a) complete absence of oxygen resulted in cessation of root growth; (b) many species can grow in an environ- ment with as little as 0.5 per cent oxygen, but only very slowly; (c) rate of supply rather than partial pressure of oxygen governs growth rate; (d) when the supply of oxygen was too low to permit normal growth, the growth rate varied inversely with the temperature; (e) at constant temperature, growth varies directly with partial pressure of oxygen; and (/) there is a critical oxygen concentration for growth, but the ex- 146 Mineral Nutrition of Plants treme differences in critical concentrations of oxygen between species did not appear to be over 1 or 2 per cent.* Another factor of considerable significance, reported by Cannon (23, 21), Bergman (8), Snow (64), Elliott (■?■?), and Loehwing (47,48), is that many plants have very few or no root hairs under conditions of low oxygen supply. There is some evidence that when roots are deprived of oxygen or brought into an environment much lower in oxygen than normal, the roots may die, then be replaced by new, more stubby roots, or the roots may stop growing for a period and then start growing again very slowly. This would seem to indicate that when brought into conditions of poor aeration, some plants may undergo a change which adapts them better to partially anaerobic conditions, providing the demands on the roots are not too great during the period of adjustment. Much addi- tional work needs to be done on this point to determine the behavior of plants when the soil on which they are growing is waterlogged either through heavy rains or improper irrigation. Went (75) reported results with tomatoes which suggest that if the plant can survive until a new root system can be established where the air supply is more favor- able, the plant can continue to grow. He obtained increased growth with tomatoes in nonaerated nutrient solutions which had sent new roots into well-aerated moss (moistened with nutrient solution) tied around the stem above the level of the solution. There is considerable evidence that roots cannot function normally to absorb water and nu- trients in the absence of oxygen. Thus, it does not appear that oxygen supplied to part of the root system of a plant would be sufficient to maintain the normal functions of the remaining nonaerated part of the root system. Instead, the old roots would die and the plant would be- come dependent upon the new root system. Boynton (16) found that few new roots were formed on apple trees when the oxygen level in the soil atmosphere fell below 15 per cent, and when oxygen was at 10 per cent or below and carbon dioxide was from 5-10 per cent, both * Cannon designated the lower critical pressure as that partial pressure at which growth ceases, and the upper critical pressure as that at which growth proceeds normally. Between these two values growth continues, but at a rate and in char- acter which are subnormal. Page and Bod man 147 root and top growth were seriously affected. Cannon (24) also reported that root growth (with corn) was below normal at oxygen concentra- tions of less than 10 per cent. Since the carbon dioxide content of the soil atmosphere increases as the oxygen supply decreases, and carbon dioxide in rather high concen- trations has an apparently toxic effect on roots, it has been suggested that the poor root growth associated with poor aeration is a result of carbon dioxide toxicity rather than lack of oxygen. Vlamis and Davis (7^) found that when barley and tomato plants were exposed to carbon dioxide, a lethal effect was produced. Chang and Loomis (27) bubbled air, nitrogen, and carbon dioxide through culture solutions and found that carbon dioxide reduced both water and nutrient uptake. They also stated that toxic to slightly toxic concentrations of 10-20 per cent carbon dioxide are probably more commonplace in soils than limiting concen- trations of 1-2 per cent oxygen. Parker (55), however, concluded that the carbon dioxide content of the soil is not important in influencing the absorption of inorganic nutrients by plants. Hoagland and Broyer (59) and Arrington and Shive (4) reached a similar conclusion for nu- trient solutions. It thus appears that the main effect of carbon dioxide in most soils is that, when present in large amounts, oxygen tends to be deficient and, through pH changes, the solubility of certain nutrients may be adversely affected. It seems amply apparent from the foregoing material that deficient oxygen in the soil would certainly limit the normal growth and exten- sion of roots and development of root hairs and, hence, affect the avail- ability of plant nutrients. Under specialized conditions, as in the green- houses, otherwise normal plants are sometimes found growing with a very small root system where the soil is kept excessively wet and large quantities of fertilizers are used. The plants appear to be dependent upon oxygen dissolved in the irrigation water; this is in a sense a solu- tion culture and in practice it is found that the plants are quite sensitive to any change in their environment. These plants can be grown with limited root systems where the soil functions chiefly as a support. It has not, however, been demonstrated that it is economical or desirable to grow plants in this way. For almost all soils, however, it appears that healthy, productive plants can only be expected when an extensive root 148 Mhieral Nutrition of Plants system is established. Such root systems are only established in soils when soil aeration is good. Loehwing (48) stated: "Improper composition of soil air manifests itself in reduced, slow-growing root systems, inadequate absorption, short-lived, discolored foliage and delay or failure of reproductive pro- cesses." Albert and Armstrong (/) found that fruit bud shedding and poor plant growth in cotton definitely resulted from poor soil aeration. Cannon (24) stated: ". . . it can be seen therefore, that there comes a point in the diminution of the oxygen content of the soil atmosphere when the growth of the root ceases because it is no longer sufficient to supply demands for energy correlated with physiological activities of higher temperatures." He stated further that ". . . in puddled soils with consequent poor aeration, and in summer, the matter of oxygen supply to the roots must be acute." He concluded that for corn to attain a fair rate of root growth at high temperature, aeration must be good indeed. PHYSICAL PROPERTIES AND NUTRIENTS The role played by soil clays in exchange reactions, and replenish- ment of the soil solution by replaceable metallic cations subsequent on their depletion by plants, is prima facie evidence of the significance of the soil particle-size distribution upon nutrient availability. This state- ment is not intended to minimize the fundamental importance of the mineralogical species present in the clay fraction, the possibility of non- replaceable fixation of ions, or the degree of accessibility to solutions and living roots of the sites of exchange reactions. There is, however, much evidence of significant positive correlation between amounts of exchangeable cations and fineness of soil texture. For illustration, reference may be made to analyses by Hosking (43) of the cation exchange capacity at pH 9 of a number of soils and their mechanical fractions from Australia and New Guinea. The soils repre- sent a wide variety of genetic groups and lithologic origins. The ex- change capacities per unit mass are by far the greatest for the material finer than 1 micron and, where observed in the coarser fractions, are, with apparent justification, attributed to incomplete dispersion. The exchange capacities of the whole soils, corrected for organic matter, increase with content of clay finer than 1 micron. This is particularly marked for those soils reported to contain some form of montmorillo- Page and Bod man 149 nite. In an earlier paper (43), Hosking expressed the opinion that un- identified amorphous inorganic materials with high exchange capacity may influence the exchange capacities of soils more often than is gen- erally supposed. Rubashov (6/) working chiefly with certain beet-producing cherno- zems, of loess and "loesslike" origin, claims that the water-stable aggre- gates (7 mm. > 0.25 mm.) present possess physico-chemical properties distinctly different from the soils from which he isolated them. The amount of "humus" present in the surface 8 inches was found to in- crease with abundance of water-stable aggregates. It also tended to increase with the coarseness of the aggregates. There was some slight indication, further, that the amount of absorbed calcium may have been greater in the more highly aggregated soils. Aggregates coarser than 0.25 mm. (macroaggregates) separated from a given soil consistently con- tained more humus and absorbed calcium than those finer than 0.25 mm. (microaggregates) obtained from the same soil. This investigator (60) reported a similar relationship for podzolized soils. Improved structural quality, according to Rubashov, is associated with higher content of total humus and of "structure-forming humin substances." The content of mobile nitrogen, acid-soluble phosphorus (Truog method), and exchangeable potassium in water-stable aggregates was found by Rubashov to be greater than in the soil as a whole. This was interpreted as indication that the water-stable aggregates strongly influence the nutritional value of a soil, as well as its physical and physico-chemical properties. Although more pronounced than those of the organic matter, the higher levels of these nutrients were associated with, and by Rubashov attributed to, the higher content and particular quality of the aggregate-forming organic matter. Amongst many other experiments on the significance of different structural conditions in the soil, Doyarenko (see 45) determined the amounts of nitrates found after six weeks in vessels of moist soil con- sisting of different aggregate size classes. Increases in nitrate formation paralleled increases in aggregate diameter and were attributed to the greater abundance of coarse ("noncapillary") pores amongst the coarse aggregates, which produced more favorable conditions for biological oxidation. Rubashov's (6/) results with respect to organic matter appear to be 150 Mineral Nutrition of Plants in general agreement with those obtained by Metzger and Hide (50) and by Weldon and Hide (74) who found, respectively, that organic carbon, and carbon and nitrogen content, were distinctly higher in well-aggregated than in poorly aggregated fractions separated from some Kansan soils. The percentages of carbon and nitrogen present, moreover, were found to decrease with a decrease in the size of aggre- gates except for the colloidal fraction, which was high in both elements. Applications of manure to alfalfa grown on one of the soils caused in- creases in the carbon content which were greater in the well-aggregated fraction. These results differ from those of Clarke (31) who examined the aggregation and total nitrogen content of certain South Australian soils and concluded that nitrogen content is not significantly related to and has no significant effect on water-stable aggregation. Breazeale and McGeorge (ij) in one of a series of investigations on soil structure made by the soils group in Arizona, examined the effect of puddling upon microbial nitrogen transformations in some alkaline- calcareous soils. Puddling was brought about by soil manipulation, by means of an electrical vibrator, at the moisture equivalent. Nitrification was found to be completely stopped by such treatment during subse- quent incubation, and denitrification became active. When nitrogen was originally added either as sodium nitrate or ammonium sulfate, a greater loss of nitrogen was observed from the puddled than from the unpuddled soil and the loss was entirely as gaseous nitrogen. Brea- zeale and McGeorge state that dry fallowing fully restores the nitri- fying power of a puddled soil. Some benefit is obtained by the addition of a dust mulch to the surface of puddled soils. This is explained by up- ward movement, when the soil is moist, of dissolved ammonia into the layer of loose soil, and also by the somewhat more friable condition developed by the slower drying of the puddled soil itself. Both effects produce a condition more favorable to microbial nitrification. In experiments upon the decomposition of organic matter in puddled and unpuddled Gila clay loam, McGeorge and Breazeale (5/) found that addition of dry and rotted alfalfa to unpuddled soils resulted in excellent growth of barley seedlings. No nitrogen deficiency was ap- parent. Seedlings on the puddled soil made poor growth, were yellow, Page and Bodman 151 and appeared to lack both nitrogen and water. Puddling very seriously interfered with root growth. In one set of cultures, after incorporation of the alfalfa which underwent anaerobic decay in the subsequently puddled soil, the soil was allowed to dry thoroughly after planting; the growth was very poor. The unfavorable soil condition prevailed for a period of three months, as evidenced by unsatisfactory development of a second crop sown after that length of time. The results were at- tributed to the persistence of toxic products of anaerobic decomposition. These authors stated that "the productivity of puddled soils may be seriously reduced by the incorporation of organic matter while the puddled condition still exists" and, "in utilizing organic matter in the rebuilding of soil structure productivity may be lost, even though the structure is regained, if proper precautions are not taken in the use of organic matter." Unsatisfactory response of tomato plants was obtained with treble superphosphate and both ammonia and nitrate nitrogen additions to puddled soils as compared with the same additions to the same soils in the unpuddled state. The effect of the puddled soil condition on nitrifi- cation (iy) has already been mentioned. In the opinion of these inves- tigators, lack of adequate aeration in the alkaline calcareous soils of Arizona may produce a deficiency in carbon dioxide with the resultant formation of normal carbonates and hydroxides. Plant growth is then affected by an unfavorable pH which produces a phosphate deficiency. In other experiments McGeorge and Breazeale examined phosphate, potassium, and calcium uptake by means of Neubauer tests with rye. It was found that soil puddling reduced the availability of these nu- trients. Interpretation by Gorbunov of his experiments with nonreplaceable potassium (j6) involves a relationship between soil structure and the availability of this element. Gorbunov estimated the amount of replace- able potassium in potassium-saturated samples of chernozem soils and a podzol clay by electrodialysis. Except for a control, the materials, after saturation and before electrodialysis, had been previously dried for 18 hours at different temperatures (40-1050 C.) and, presumably, with different degrees of completeness thereby, before re wetting and maintaining wet for 72 hours. The electrodialysis which followed 152 Mineral Nutrition of Plants brought into solution, after about 2 to 3 hours and during the next 15 to 20 hour period, an amount of potassium which, in all cases, was greatest for the undried materials and which diminished as the tem- perature of drying increased. In explanation, Gorbunov regarded the heating and drying as re- sponsible for a process of aging of the colloidal material during which much of the diffuse part of the electrical double layer was destroyed. That is, combination of potassium ions with residual negative charges on the absorptive material made the latter less electronegative. He ap- pears to have regarded this change as in large part irreversible and accompanied by the formation of microaggregates with reduced dis- persibility. According to this hypothesis, aggregation may hinder the replaceability of the ion concerned, in this case potassium. Gorbunov made no statements concerning the size of the "microaggregates." On the other hand, he considered that organic matter plays an important role in the fixation process. Gorbunov's hypothesis does not satisfac- torily explain the relationships reported by Rubashov (61) between aggregate size and replaceable potassium. Martin, Overstreet, and Hoagland (49), on the other hand, found that potassium fixation may be considerable in moist soils and, contrary to Gorbunov, do not con- sider that drying is directly involved. The clarification of these apparent inconsistencies awaits more work by chemists and physicists on the mechanisms of the aggregation process. Experiments were recently conducted by Greacen (57) upon the applicability of Boyd's equations (75) to cation exchange in columns of soil aggregates. Greacen gave special attention to mechanisms which might control rates of exchange between soil-adsorbed cations and cations present in dilute, permeating solutions. Rates at which soil aggregates of different sizes released their cations were measured by passing solutions at controlled velocities through very short columns of aggregates. Cation exchange for the Yolo and Aiken soil aggregates within the "fine" size class, 0.3-0.1 mm., was found to be an equilibrium process during passage, through a column of calcium aggregates, of N/50 mag- nesium chloride solutions at macroscopic velocities of less than 0.3 cm. per second. For coarser aggregates, > 0.6 mm., of Aiken soil, diffusion Page and Bod man 153 of ions through the individual aggregates appeared to control the rate of exchange, for macroscopic velocities in excess of 0.004 cm- Per second. The two soils did not behave identically and, for soils in general, much individuality of behavior seems probable. Rates of internal diffusion were found to limit the exchange rates with the fine aggregates also, provided that the macroscopic velocity of the permeating fluid was made sufficiently great (e.g., 1-2 cm./sec). Such high velocity, however, is unlikely to be of direct interest in prob- lems of plant growth in soils. From these experiments it appears that ionic diffusion rates within aggregates may possibly affect the ready availability of exchangeable cations, particularly if the aggregates are large, impenetrable by roots, and not easily dispersible. ROLE OF SOIL AERATION IN NUTRIENT UPTAKE Soil aeration not only affects the extent and character of root growth, but is also of extreme importance through the effect oxygen has upon the assimilation of nutrients by roots. Much of the more conclusive evi- dence again is to be found from studies in solution cultures, primarily with excised root systems. Since this work will be the topic for another paper on this symposium, it will not be reviewed in detail here, but a few papers will be cited to indicate the essentiality of oxygen in absorp- tion. The work of Steward and his colleagues (65) showed that, for tissues, aerobic respiration supplies the energy necessary for salt absorp- tion against a concentration gradient. The extensive studies of Hoag- land and Broyer ( 39, 40) proved that salt accumulation by roots is also dependent upon aerobic metabolism, and oxygen is one of the indis- pensable requirements for salt accumulation (movement of salt against a gradient) by excised barley roots. These studies were made on assimi- lation of potassium, halide, and nitrate. They also showed that a rela- tively high concentration of carbon dioxide is required to greatly de- press salt accumulation. Where oxygen had been carefully excluded, they did not observe accumulation of salt against a gradient. The recent work on this problem is excellently summarized by Hoagland (41). In work with growing tomato plants in solution cultures, Arnon and Hoagland (_?) found that roughly 1.4 times as much each of potassium, 154 Mineral Nutrition of Plants nitrate, phosphate, calcium, and magnesium was taken up from aerated solutions as from nonaerated solutions. The size of plants and yield of fruits were correspondingly high. Arrington and Shive (4) found that absorption of ammonia nitrogen was 1.34, and of nitrate nitrogen 1.30, times greater from aerated than from nonaerated solutions. Pepkowitz and Shive (56) found that absorption of calcium and phosphorus was directly dependent upon dissolved oxygen supply, whereas absorption of potassium was not so materially influenced. Chang and Loomis {27), on the other hand, found that aeration increased absorption of potas- sium most, followed, in order, by nitrogen, phosphorus, calcium, and magnesium. Direct evidence for the effect of soil aeration on nutrient availability is scarce, but the evidence available is significant and highly suggestive. Probably the main reason why there are few data on this subject is that there is no good way of measuring or characterizing soil aeration; thus, it has been difficult to prove that certain conditions observed in the field have been caused by poor soil aeration. Many attempts have been made to measure the partial pressure of oxygen in the soil atmosphere di- rectly, but little success has been attained. One difficulty has been that the analytical methods available required a rather large gas sample. This was not easy to extract from the soil in a condition representative of the atmosphere surrounding plant roots. Another difficulty is that measurement of oxygen in the soil air may not indicate the amount of oxygen dissolved in the water bathing individual roots, for normal soil changes in the oxygen concentration of the soil atmosphere may occur too rapidly for equilibrium between dissolved and gaseous oxygen to be maintained. With the recent introduction of the Pauling oxygen analyzei, which requires only a few cubic centimeters of gas and which indicates partial pressure of oxygen directly and automatically, a new tool is available which should give valuable information concerning the mechanism of gas interchange in the soil. Even without data on the actual aeration status of the soil, a few ex- periments and observations have been reported which emphasize the importance of aeration to normal plant growth and the effect of inade- quate soil aeration upon nutrient availability. Page and Willard (54) observed in an experiment which compared different tillage methods for corn, that marked potassium deficiency symptoms occurred in corn Page and Bod man 155 on plots which had been prepared by disking only. Milder deficiency symptoms were observed on these same plots even where 300 pounds of 0-14-7 had been placed in the row at planting time. The air space porosity in these plots was only 14.2 per cent as compared with 26.9 per cent for adjacent plowed plots, which showed no potassium de- ficiency symptoms. These two adjacent plots were comparable except for the one variable of tillage in preparing seedbeds. Bower, Browning, and Norton (14) obtained symptoms of both nit- rogen and potassium deficiencies in the first reported experiment which demonstrated that tillage had a direct effect upon availability of nu- trients. They observed definite nitrogen deficiency symptoms on corn which was growing on land prepared by disking or subsurface tillage, even where nitrogen fertilizer was applied. Corn on plowed land showed almost no evidence of nitrogen deficiency. Similar results were obtained with potassium, but the differences were not quite so clear cut. Where no fertilizer was applied, however, the potassium content of plowed corn was 70 per cent higher than the average of corn grown by the other three treatments. Significantly, they found no difference in exchangeable potassium content of the soils between these plots, yet the potassium was definitely more available to the corn plant from the plowed plots. They indicated that tests showed more ferrous iron (indi- cating lack of aeration or reducing conditions) in the soils of those plots which were not plowed and on which potassium deficiency was observed. Lawton (46) made a study of the effect of aeration on absorption of nutrients in pots in the greenhouse, where different degrees of com- paction, different water levels, or forced aeration could be maintained. He found that where pore space was reduced, either by compaction or increase in water content, absorption of potassium was much less than from a normal soil. Under these conditions root growth was also seri- ously reduced. Forced aeration eliminated the distinct potassium de- ficiencies, even though air was forced through the soil for only 30 min- utes a day. He observed that absorption of potassium is more dependent upon soil aeration than is uptake of nitrogen, calcium, magnesium, or phosphorus. This is in agreement with the findings of several investi- gators working with solution cultures. Smith and Cook (6j) studied the effect of soil aeration and compac- 156 Mineral Nutrition of Plants tion on nitrification and the growth of sugar beets. They found that in every case compaction decreased beet yields, and artificial aeration tended to overcome the harmful effects of compaction, indicating that low aeration was limiting yields. They also found that compacted soil, in which large amounts of nitrogenous material had been incorporated, had been prevented from rendering these nitrogenous substances avail- able as nitrates, at least to a level comparable with that which might normally be expected and which was demonstrated on the same soil that had not been compacted. Trogdon (68) found in a study of effectiveness of various nitrogen fertilizers that maximum utilization of added fertilizer elements was not obtained if air space porosity was low. Wherever soil porosities were low, tests showed the presence of reducing conditions in the soil throughout much of the two rather wet seasons and definite phos- phorus and nitrate deficiency symptoms were observed even where 700 pounds of ammonium sulfate and 800 pounds of 0-10-10 fertilizer had been applied. It is, of course, well known that oxygen is required for the produc- tion of nitrates, for the fixation of nitrogen, and, in fact, for the con- tinued activity of most of the important types of microorganisms in the soil. A fact which is not so widely considered, however, is that the requirement for oxygen by microorganisms which are actively decom- posing organic matter in the soil is comparatively high. It appears very likely, in fact, that the commonly observed depression in plant growth immediately after adding large amounts of green manure or other organic matter to a soil may be due more to competition between the higher plants and the microorganisms for the limited supply of oxygen than to any other single factor. The full significance of this point re- mains to be worked out, but it is worthy of further study. Trogdon (68) found that, where large quantities of fresh organic matter were added to soil in contact with fertilizer banded below the surface, there was a reduction in yield as compared to no organic matter. This was apparently associated with deficient oxygen supplies since definite nit- rogen, phosphorus, and potassium deficiency symptoms were observed in the plants and reducing conditions could be demonstrated in the soil near the organic matter. It is likely that some of the peculiar results Page and Bodman 157 sometimes observed in glass-house vegetable production can be ex- plained as due to temporary partial reducing conditions brought about by turning under large amounts of organic matter where temperature and moisture conditions favor rapid decomposition. Where reducing conditions are developed in the soil either through additions of large amounts of organic matter, through waterlogging, or compaction, or a combination of these factors, it is, of course, well known that important chemical changes are produced in the system. Under reducing conditions, carbon dioxide is usually reduced to me- thane; nitrates to nitrites, ammonia, or free nitrogen; sulfates to hydro- gen sulfide; ferric to ferrous iron; and trivalent to divalent manganese. There is considerable evidence that most of the reduction is brought about by microorganisms and that the presence of readily decomposable organic matter greatly favors the development of reducing conditions wherever oxygen supply might be low. In most soils, however, every effort is, or should be, made to prevent the occurrence of extreme re- ducing conditions, so that too great emphasis on the nature of the chemical status of the reduced soil is probably not warranted for this discussion. Of more significance is the effect of even a partial reduction in the supply of oxygen available to plant roots, caused either through utiliza- tion of available oxygen by microorganisms, or through reduction in the rate of supply through the soil pores. Since oxygen is necessary for normal root functioning and absorption of water and nutrients, any process which significantly reduces the supply will have an adverse effect on root growth and availability of nutrient elements. MECHANICS OF SOIL AERATION The work reported in the preceding sections indicates that nutrient uptake and availability, particularly of potassium, is affected by soil aeration. As was mentioned earlier, the detailed mechanics of soil aera- tion are not well understood, but workers in the field are quite well agreed that practically all aeration occurring in the soil results from gase- ous interchange by diffusion. Oxygen diffuses into the soil and carbon dioxide diffuses out, simultaneously. Factors such as wind, barometric pressure changes, or flushing by light rains or irrigations, apparently 158 Mineral Nutrition of Plants contribute little to the movement of soil gases. In the past, attempts at studying soil aeration have been directed mostly toward studying partial pressures of oxygen in the soil air, by measuring porosity (usually at a standard moisture content) or by studying permeability of soil to gases under a pressure differential. Each of these methods has distinct shortcomings. Difficulties in measuring oxygen have already been mentioned, but such work should be continued making use of the newly-available oxygen analyzers. Determination of porosity gives at best an indirect measure of aeration of a soil, particularly if all measurements are made at some standard moisture condition, since this exact moisture condition may only occur in the field rarely and, hence, cannot indicate the aeration conditions under which the plants are living. An improvement in this measurement is made possible by measuring porosity at field moisture with the pressure pycnometer (52). With this instrument it is possible to measure the pore space actually filled with air at the moisture condition prevailing in the soil at the time of sampling. Another shortcoming of porosity measurements is that the effectiveness of air pore space in the soil is limited by the de- gree to which there are open channels or continuous pores to and through the surface so that free diffusion can occur. It is easy to see how a soil with apparently very favorable porosity in the root zone could be quite unsatisfactory for root growth if the surface were sealed through formation of a crust. Possibly the most important benefit of a mulch comes through its action in maintaining porosity unblocked or sealed through to the surface, so that free diffusion into the soil can occur. The chief drawback to characterizing aeration in terms of air permeability is that gas flowing under a pressure differential may be subject to such factors as turbulent flow, friction, and others which are not of significance in free diffusion. In recent work, Blake (9, to) has studied the diffusion process. He found, as have previous workers, that there is a linear relationship be- tween diffusion and porosity, but, unlike most work reported in the literature, he did not find a constant factor to express the relationship between porosity and diffusion. The coefficient was different for soils of different properties. Thus, two soils might have identical porosities when expressed on a volume basis, but, if they differ in the character of Page and Bod man 159 the pores or in the arrangement of the pore spaces, they might behave quite differently in so far as diffusion, and hence aeration, is concerned. One soil having 30 per cent air space porosity with interconnected channels open to the atmosphere would certainly be better aerated than another having the same percentage porosity but having few interconnected channels. It was suggested, in fact, that measuring the diffusion characteristics of a soil and evaluating porosity in terms of the diffusion process rather than on a strict volume percentage basis offers a most promising method of attack on the difficult problems in soil aeration. To accomplish this and make an even more direct determination, a new method of measurement was devised. This procedure has not yet been described in the literature, but it appears to have considerable promise and value as a research tool. Briefly, it consists of introducing a large excess of carbon dioxide into a soil from a cylinder of gas (the gas is introduced through a metal capillary tube arranged into a pointed sampler) to sweep out soil gases from a rather large volume of the soil. The amount of gas is not critical as long as an excess is used. When the flow of gas has been stopped, the oxygen analyzer is attached to the same tube and small samples are withdrawn as frequently as de- sired. By this means it is possible to measure the rate at which oxygen diffuses into the soil to dilute the carbon dioxide until normal soil atmosphere concentration is approached. The results are expressed as a time curve showing renewal rate for diffusion of oxygen into the soil. The method is simple and direct: no foreign materials are introduced, diffusion proceeds unhampered through natural soil pores at existing field moisture content, the results are easily interpreted, and the samp- ling site is not changed, so that measurements can be repeated at the same location as often as desired throughout a whole growing season. A similar procedure which uses standard 3-inch cores of soil in natural structure has also been used in the laboratory. Even with a limited number of determinations, some interesting data were obtained with the techniques described. Thus, renewal rate — the time for oxygen to reach the normal concentration usually found in the soil — was two hours in a freshly plowed soil and three hours in a corn plot (determinations made in late fall) ; at four hours the oxygen level 160 Mineral Nutrition of Plants had reached only 16 per cent (instead of the usual 18.5-19.5 per cent) in an alfalfa-brome-grass plot, and in continuous bluegrass plot the oxygen level had risen only to 14 per cent in 22 hours. Temperatures were low, and it was not thought that production of carbon dioxide in the soil was a significant factor, although it is a remote possibility in these experi- ments. It is surprising to note that renewal rates are considerably lower than published estimates have suggested. Direct determinations of oxygen in the soil atmosphere with the oxygen meter have shown very few oxygen concentrations lower than 18.5 per cent. Even on experimental plots (52, 55) where soil structure was known to be poor and where crop yields were definitely limited by poor aeration, measured oxygen levels were comparatively high. It was concluded that on most soils aeration is adequate during much of the season when the soil water is at or below field capacity. At times of rains or heavy irrigations soils of good structure drain readily, and normal aeration is re-established before the supply of dissolved oxygen in the soil water is exhausted by the plant. In soils of poor structure, however, aeration may not be so rapidly re-established, and, if the oxygen supply runs out and is deficient for any length of time, the plant may undergo a shock or change which will adversely afTect its future growth and development. Much work needs to be done on the response of plants to intermittent periods of low oxygen supply, to de- termine the extent to which root growth and absorption of nutrients are affected. As an example of a situation in which plant growth is definitely limited by physical properties, and to show that these properties are subject to modification, reference will be made to the soil structure rotations of the Ohio Station (§2, 5.?). Various rotations were estab- lished in 1936 on Paulding clay to compare the effects of different cropping systems on soil structure. Summarized results are presented in Table II. It can be seen that "structure building" rotations which in- clude sod crops have effected an improvement in productive capacity of the soil while, with those rotations which do not provide for adequate organic matter return, productivity has sharply declined. The changes described have been brought about in a comparatively short time. The effect of fertilizers has been studied on the same plots Page and Bodman 161 but no consistent or large increases have been obtained; in fact, in some years there was a loss in yield where fertilizers were added. Table III shows the average corn yields and gains from fertilizers in recent years. It is significant to note that the natural fertility status of these soils is high: yields as high as no bushels of corn have been obtained when the season and the physical condition of the soil were both favorable. TABLE II Relative Average Unfertilized Corn Yields for Different Rotations (Paulding, Ohio) (Actual 1936 Yield = 100 per cent) Rotations 1945 1942-1948 per cent per cent 4 year (corn, oats, 2 yr. alfalfa or alfalfa-brome grass) 191 147 3 year (corn, oats, alfalfa or sweet clover) 147 106 2 year (corn, oats) 90 71 Continuous corn 61 36 Note: 1945 best corn year during period. Data before 1942 not used because rotations had not been well established before that year (55). TABLE III Average Unfertilized Corn Yields and Response to Fertilizer Applications (Paulding, Ohio) Av. Gain Yield for 1945,1946, Fer- Rotations 1948 tilizer bu. bu. per acre per acre 4 yr. (corn, oats, 2 yr. alfalfa or alfalfa-brome grass) 62.5 3.9 3 yr. (corn, oats, alfalfa or sweet clover) 56 . 1 1.4 2 yr. (corn, oats) 36.6 8.7 Continuous corn 16.3 8.2 Note: 1947 yields not included since yield data were not taken for fertilized vs. nonfei tilized plots. Fertilizer schedule: 150 lb. 0-20-20 /acre/yr. plus 300 lb. 20-0-0 before corn on last two rotations. ^ # y 1 62 Mineral Nutrition of Plants ^'i.si-"A*:^&«» BHH^^^HHHKSNHI Figure i. Appearance of structure rotation plots on July 20, 1949, Pauld- ing, Ohio. Upper: corn, oats, two-year alfalfa rotation. Lower: continuous corn. Poor drainage characteristics and poor soil structure, which have de- veloped since 1936, definitely limit plant growth through poor soil aeration on this plot; there has been a definite improvement in soil structure and, hence, in soil productivity on the plot shown at the top. Page and Bod matt 163 These yields were obtained in areas which had received no fertilizers for several years. Yields as low as 5 bushels have been obtained from plots where soil structure has been allowed to deteriorate. Fertilizer application at the rate shown in Table III raised the yield on the cor- responding plot in that year to only 1 1 bushels. The soils under different rotations show very great differences in physical properties, especially as regards drainage. The conditions of soil and crops shown in Figure 1 and the measured physical properties including porosity, aggregation, and diffusion all correlate well with yields, indicating that the yields are definitely limited by poor soil aeration. On the continuous corn plots the corn is very poor and shows many deficiency symptoms for the various nutrients even though these plots regularly receive 150 pounds per acre of 0-20-20 and 300 pounds of 20-0-0. In this case physical properties of a soil definitely limit the growth of the plants and the over-all availability of the plant nutrients, and addition of fertilizers does not raise the yield to a level where farming would be profitable. These studies furnish a striking example of the inadequacy of a strictly chemical approach in studies on crop production and emphasize the close interrelation between all plant growth factors, physical as well as chemical. To summarize, oxygen or proper soil aeration is definitely needed for normal root growth, and for normal absorption of nutrients by roots. Nutrients may be abundant and available as shown by chemical tests, but largely unavailable to plants if the soil air supply is not adequate. Soil structure is the governing property which not only controls air supply but also affects the growth of roots as they penetrate the soil. Large amounts of organic matter, added for the purpose of improving soil structure, may cause temporarily harmful rather than beneficial ef- fects through production of reducing conditions in the soil. Thus, soil structure must be taken into account in interpreting the results of fer- tilizer trials or applications, and at the same time it must be realized that soils can be improved not alone through addition of fertilizers but through improvement of their physical properties as well. 164 Mineral Nutrition of Plants REFERENCES 1. Albert, W. B., and Armstrong, O., Plant Physiol., 6:585 (1931). 2. Allison, R. V., and Shive, J. W., Am. f. Botany, 10:554 (1923). 3. Arnon, D. I., and Hoagland, D. R., Soil Sci., 50:463 (1940). 4. Arrington, L. B., and Shive, J. W., Soil Sci., 42:341 (1936). 5. Baicourt, A. W., and Allen, R. C, Proc. Am. Soc. Hort. Sci., 39:423 (1941). 6. Baver, L. D., and Farnsworth, R. B., Soil Sci. Soc. Am. Proc, 5:45 (1940). 7. Baver, L. D., Soil Set'., 68:1 (1949). 8. Bergman, H. J., Ann. Botany, 34: 13 (1920). 9. Blake, G. R. "Aeration of soils with special emphasis on the diffusion process in soil air removal." Ph.D. Diss., Ohio State University (1949). 10. , and Page, J. B., Soil Sci. Soc. Am. Proc, 13:37-43 (1948). 11. Bodman, G. B., Bull. Am. Soil Survey Assoc, 12:111 (1931). 12. , and Day, P. R., Soil Sci., 55:225 (1943). 13. , and Rubin, }., Soil Sci. Soc. Am. Proc, 13:27-37 (1948). 14. Bower, C. A., Browning, G. M., and Norton, R. A., Soil Sci. Soc Am. Proc, 9:142 (1944). 15. Boyd, G. E., Myers, L. S., and Adamson, A. W., /. Am. C/iem. Soc, 69:2849 (1947). 16. Boynton, D., Proc Am. Soc. Hort. Sci., 37:19 (1939). 17. Breazeale, J. F., and McGeorge, W. T., Ariz. Agr. Expt. Sta. Tech. Bull. 6g (1937). 18. Bryant, A. E., Plant Physiol., 9:389 (1934). 19. Buehrer, T. F., and Rose, M. S., Ariz. Agr. Expt. Sta. Tech. Bull. 100 (1943). 20. Bushnell, J., /. Am. Soc. Agron., 27:252 (1935). 21. Cannon, W. A., Carnegie Inst. Wash. Yearbook, 24:289 (1924-25). 22. , Carnegie Inst. Wash. Yearbook, 15:74 (1916); 16:82 (1917); 17:81 (1918); and 31:294 (1932). 23. , Carnegie Inst. Wash. Publ. j68 (1925). 24. , Science, 58:331 (1923). 25. , and Free, E. E., Carnegie Inst. Wash. Publ. 368 (1925). 26. , and Free, E. E., Science, 45:178 (191 7). 27. Chang, H. T., and Loomis, W. E., Plant Physiol., 20:221 (1945). 28. Clark, H. E., and Shive, J. W., Soil Sci., 34:37 (1932). 29. Clark, Cyrus L., Soil Sci., 35:291 (1933). 30. Clarke, G. B., and Marshall, T. J., /. Council Sci. Ind. Research (Australia), 20:162 (1947). 31. Clarke, G. B., /. Council Sci. Ind. Research (Australia), 21:51 (1948). Page and Bod man 165 32. Day, P. R. "The moisture potential of soils by the cryoscopic method." Ph.D. Diss., Berkeley, California (1941). 33. Elliot, G. R. B., Ecology, 5:175 (1924). 34. Erickson, L. E., Am. J. Botany, 33:551 (1946). 35. Gilbert, S. F., and Shive, J. W., Soil Sri., 53:143 (1942). 36. Gorbunov, N. E., Khimizatsiya Sotsialist. Zemledeliya, No. 2-3:82 (i936)- 37. Greacen, E. L. "Applicability of Boyd's exchange equations to cation exchange in soil columns." Ph.D. Diss., University of California, Berke- ley, California (1949). 38. Hoagland, D. R., and Martin, J. C, Soil Sci., 36:1 (1933). 39. Hoagland, D. R., and Broyer, T. C, Plant Physiol., 11:471 (1936). 40. Hoagland, D. R., and Broyer, T. C, /. Gen. Physiol., 25:865 (1942). 41. Hoagland, D. R., Lectures on the inorganic nutrition of plants (Chronica Botanica Co., 1944). 42. Hoffer, G. N., Better Crops With Plant Food, 29, no. 1:6 (1945). 43. Hosking, J. S., /. Council Sci. lnd. Research {Australia), 21:38 (1948). 44. Knight, R. C, Ann. Botany, 38:305 (1924). 45. Krause, M., Landw. Jahrb. 73:603 (1931). 46. Lawton, Kirk, Soil Sci. Soc. Am. Proc, 10:263 (1945). 47. Loehwing, W. F., Proc. Iowa Acad. Sci., 38:71 (1931). 48. , Plant Physiol., 9:567 (1934). 49. Martin, J. C, Overstreet, R., and Hoagland, D. R., Soil Sci. Soc. Am. Proc, 10:94 (I945)- 50. Metzger, W. H., and Hide, }. C, /. Am. Soc. Agron., 30:833 (1938). 51. McGeorge, W. T., and Breazeale, J. F., Ariz. Agr. Expt. Sta. Tech. Bull. 72 (1938). 52. Page, }. B., Soil Sri. Soc. Am. Proc, 12:31 (1947). 53. , and Willard, C. J., Soil Sci. Soc. Am. Proc, 11:81 (1946). 54. , and Willard, C. )., Soil Sci. Soc. Am. Proc, 11:77 (I94^)* 55. Parker, F. W., Soil Sci., 20:39 (1925). 56. Pepkowitz, L. P., and Shive, J. W., Soil Sci., 57:143 (1944). 57. Richards, L. A., and Weaver, L. R., Soil Sci., 56:331 (1943). 58. Richards, L. A., Soil Sci., 68:95 (I949)- 59. Rubin, J. "The influence of externally applied stresses upon the struc- ture of confined soil material." Ph.D. Diss., Berkeley, California (1949). 60. Rubashov, A. B., (Russian). Pochvovedenie, No. 4 (April 1940). 61. , (Russian), Pochvovedenie, No. 3:130 (March 1949). 62. Schuylenborgh, J. van, "A study on soil structure." Ph.D. Diss., Wageningen (1947). 63. Smith, F. W., and Cook, R. L., Soil Sci. Soc. Am. Proc, 11:403 (1946). 64. Snow, L. M., Botan. Gaz., 40:12 (1905). 166 Mineral Nutrition of Plants 65. Steward, F. C, and Preston, G., Plant Physiol., 15:23 (1940). 66. Stiles, W., and Jorgensen, I., New Phytologist, 16:181 (191 7). 67. Thomas, M. D., Soil Sci., 25:409 (1928). 68. Trogdon, W. D. "The effect of acid and base forming nitrogenous fertilizers on certain soil factors affecting soil productivity." Ph.D. Diss., Ohio State Univ. ( 1949). 69. Veihmeyer, F. J., and Hendrickson, A. H., Plant Physiol., 20:517 (r945)- 70. , and Hendrickson, A. H., Soil Sci., 62:451 (1946). 71. , and Hendrickson, A. H., Soil Sci., 65:487 (1948). 72. , and Hendrickson, A. H., Soil Sci., 68:75 (1949). 73. Vlamis, J., and Davis, A. R., Plant Physiol., 19:33 (1944). 74. Weldon, Thomas A., and Hide, }. C, Soil Sci., 54:343 (1942). 75. Went, F. W., Plant Physiol.. 18:51 (1943). CHAPTER / Role of Soil Microorgan- isms in Nutrient Availa- bility A. G. NORMAN S 'oil microbiology is a field of endeavor about which considerable diversity of opinion has been expressed by workers whose interests lie in the study of factors involved in plant growth. Many com- pletely ignore the soil microbial population, apparently because it has been established by nutrient culture studies that, under certain peculiar circumstances not normally achieved in soil, the presence of micro- organisms is not essential for optimum plant growth. Others refer to the soil population only when unsatisfactory plant growth is obtained; it is then convenient to be able to ascribe the growth deficiency to "immobilization" of this or that essential nutrient by the soil micro- organisms. To a degree the views of some of the earlier soil microbiologists may have contributed to this attitude. Particular groups of microorganisms, studied in pure culture, were found to bring about reactions the product or products of which in soil might advantageously afTect plant growth. These were "beneficial" organisms. Others, capable of effecting trans- formations that might result in reduction in solubility, availability, or loss of some nutrient, were "harmful" organisms. The great bulk of the population not having obvious specialist functions was largely passed over. It was implied that in some way or other the role of the soil popu- lation with respect to plant growth could be expressed by a summation of the positive (beneficial) and negative (harmful) components. In this review consideration will be given only to the activities of microorganisms that afTect the supply of major and minor plans nu- 168 Mineral Nutrition of Plants trients in soils. Direct effects other than nutritional, or indirect effects such as those that relate to soil aggregation, will not be discussed. THE NATURE OF THE SOIL POPULATION As will be developed later, the vegetation supported by the soil and the microbial population within the soil each exert influences on the other. The relationship is peculiarly complex and cannot be readily described or denned in a single term. The activities and requirements of the plants mesh with the activities and requirements of the micro- organisms at many points. At some points these two populations may be competitive; at others the activities of the one may help to satisfy some requirement of the other; elsewhere they may be compatible or inert. As Truog has pointed out, the "living phase" of the soil, its mi- crobial population, is diverse in character. Many of the specialist or- ganisms in soil, some autotrophs and some heterotrophs, were dis- covered fairly early in the development of soil microbiology. Because almost all of these were bacteria, the impression has developed that in so far as the growth of plants is concerned, the bacterial flora of soil is the significant flora and that other groups of microorganisms, being more or less inert in their relationships to plant growth, can be safely ignored. Soil microbiology, however, is certainly not a subfield of bacteriology. Plating experiments and pseudoquantitative ecological studies on the soil population undoubtedly show that the bacteria are numerically dominant. They are, however, individually small in size and may not, therefore, occupy so dominant a position with respect to the total mass of the population as their profusion might suggest. This may particularly be the case in soils or horizons of soils in which there have been recent additions of residues of crops or other vegetation. Filamentous forms — fungi and actinomycetes — may then in mass con- stitute a major part of the active population. The available quantitative information about such forms is of dubious value. Plate counts of or- ganisms with an indeterminate habit of growth are largely meaningless because they represent as individuals both spores and hyphal fragments. In certain soils, and particularly those of lacustrine origin, algae may not infrequently be quite numerous. Protozoan forms can often be A. G. Norman 169 seen by direct microscopic examination of soil and a large number of genera of protozoa have been isolated from soil. The role of these two groups in the composite activity of the complex microbial population of which they are a part is completely obscure, and, since the abandon- ment of the "protozoan" theory of soil fertility and the censuslike studies almost twenty-five years ago, both have been largely ignored. A soil is a product of its environment and the populations which develop within the various horizons of a soil are differentiating charac- teristics no less valid than those exhibited by the organic or inorganic components of the soil. They are, however, far less readily determined or simply described and, moreover, they are dynamic and in a sense easily mutable. Particularly is this the case in cropped soils. In virgin soils or under permanent sod there is no doubt a greater degree of population stability. Soil is a medium in which many types of organisms appear to be able to retain viability for long periods though inactive. This is as true of vegetative cells as it is of bacterial endospores or the spores of fungi or actinomycetes. Perhaps the retention of water by clay surfaces provides a shell that is protective to organisms in contact with the clay. There is no ready way of determining whether an organism isolated from soil was active immediately prior to isolation or was in an inactive, dormant condition. The picture of any soil population that can be obtained is a highly indiscriminate one, and accordingly it is not easy to ascertain the ecological responses to cropping or even to major changes in the physical, chemical, or nutritional status of a soil. Even uncropped or untilled soil does not provide a biologically stable environment. Even though the carbonaceous energy supply in such soil is only occasionally supplemented, the physical environment is far from constant. There are seasonal and diurnal temperature changes, and, perhaps more importantly, there are the continually occurring erratic fluctuations in moisture content imposed by rainfall and evapora- tion. In the higher moisture range, particularly in fine textured soils, gas diffusion may be impeded, which results in alteration of the com- position of the soil atmosphere and, in the extreme case of waterlogging, in the establishment of reducing conditions and the dominance of facultative and anaerobic bacteria. 170 Mineral Nutrition of Plants THE NUTRITIONAL REQUIREMENTS OF SOIL ORGANISMS In examining the question as to the effects which soil organisms may have on the availability and supply of plant nutrients, some considera- tion must first be given to the comparative nutritional requirements of microorganisms and plants in order to determine whether they over- lap and, if so, to what extent and in what circumstances they may be competitive. Plants in general and crop plants in particular, despite all their diversities of form and structure, probably have a greater similarity of requirements than do the heterogenous forms of life that constitute the microbial population of the soil. Plants and microorganisms proba- bly require the same major nutrient elements, though no doubt wide quantitative differences occur in the levels of poverty adjustment and luxury consumption. The qualitative and quantitative range of es- sential minor elements is less well established, particularly for micro- organisms. Plants, in the language of the microbiologist, are photoau- totrophic organisms; the soil population on the other hand is predomi- nantly heterotrophic, though there are some important groups of chemoautotrophs, and some photoautotrophs that probably live hetero- trophically. The plant can and normally does satisfy its nutrient re- quirements from inorganic sources. Many heterotrophic organisms also can satisfy their nutrient requirements exclusively from inorganic sources provided that an organic source of carbon and energy is availa- ble, so, in this respect, these organisms may be said to be capable of competing with plants for the available supply of mineral nutrients. Many heterotrophic organisms, however, are able to utilize organic sources of nitrogen, phosphorus, sulfur, and, no doubt, other elements. In general these are not utilized directly but only after hydrolysis or other degradative transformation that may result in the liberation of the nutrient element in an inorganic form. Some soil organisms re- quire an organic nitrogen source, usually some simple amino acid, and, indeed, attempts have been made to use nitrogen nutrition as an aid in characterizing different soil populations (16). There is evidence that certain legumes under sterile conditions can directly utilize some soluble organic sources of nitrogen, such as aspartic acid (75) ; apparent utiliza- tion of organic nitrogen sources by nonlegumes has also been reported, A. G. Norman 171 but the maintenance of a condition of asepsis that would rule out the intervention of organisms on the surfaces of the roots has not been indubitably established. There are reserves of phosphorus and sulfur in soils in forms other than organic, so the fact that organic sources of these elements may be available to microorganisms but not to higher plants is not of special significance. There is, however, no reserve of nitrogen in soils other than that in the organic form. It might, there- fore, be supposed that the soil population could successfully compete with the plant for nitrogen and that the supply to the plant would be limited thereby. This is rarely the case, however, as will be discussed later. In so far as the mechanism of uptake of nutrient elements is con- cerned, it would seem that the process in plants and microorganisms is essentially similar. The clay fraction of soil is now recognized as being of great significance in plant growth. The physical and nutritional characteristics of a soil are determined to a large degree by the physico- chemical properties of the clay components. The organic fractions, though constituting a vital and dynamic reserve of certain nutrients, modify but cannot fundamentally alter the character of a soil imposed on it by the presence of particular clay minerals. Microorganisms in close contact with the clay colloids can no doubt compete with the clay for cations in the soil solution and by contact exchange can accept ions adsorbed on the clay. Certain soil organisms have indeed been shown to make better growth in a clay-containing medium than in conven- tional mineral nutrient media (S). Although this may not be due ex- clusively to the form of presentation and means of transfer of ions, nevertheless it appears probable that in soil, microorganisms absorb ions by a process similar to that which holds for plant roots, and in this respect they can be considered to be competitive for exchangeable ions. The cation adsorptive capacity of some bacteria has been shown experimentally to be several times that of an equal mass of barley roots (7). This observation may not be of particular significance because not all retentive positions have to be occupied by cations for growth to occur. Viability may be retained even though all cations are replaced by hydrogen or methylene blue. It appears likely, then, that bacteria and plant roots in the same absorption zone are in equilibrium with 172 Mineral Nutrition of Plants each other and with inorganic or organic colloids with respect to cation distribution. EFFECTS OF THE CROP ON THE SOIL POPULATION Although the main topic of this paper is the relationship between soil microorganisms and plant nutrition, it is pertinent to inquire just what effects may be impressed upon the microbial population of the soil by the presence of a crop. In other words, this amounts to an inquiry as to whether the population of an uncropped soil would be modified in any way by the presence of a crop. All techniques of study show that in the immediate vicinity of the root hairs, rootlets, and roots there is a great concentration of organisms, primarily bacteria, and that the root system of plants is, therefore, virtually encompassed by a zone in which microbial activities are presumably intense and are not neces- sarily identical with those occurring in soil more remote from the roots (2, 4). The biochemical activities of the rhizosphere population are somewhat obscure and yet highly pertinent to the general topic under consideration. It seems to be established that the population of the rhizosphere is not merely a more concentrated and more active version of that to be found in the soil away from the roots. Qualitative as well as quantitative differences are found. Bacteria greatly predominate over fungi or actinomycetes; gram negative rods constitute the bulk of the population; sporeformers are few. The high concentration of organisms in this zone is particularly striking. Although there is great variation in figures reported by different workers, according to particular cir- cumstances, the rhizosphere contains 5-50 times as many bacteria countable on plates as are found outside this zone. The question might therefore be asked as to whether there is much point to the study of the general soil population; perhaps the rhizosphere population is the only significant one in so far as the growth of the plant is concerned. Such an answer would ignore the fact that the root zone constitutes only a fraction of the soil under any crop other than a permanent sod. It would ignore those biochemical activities that precede the establish- ment of a particular crop on the land and its invasion by roots. It would ignore the fact that no new organisms are ordinarily introduced when a crop is planted. It would ignore the fact that when a crop is A. G. Norman 173 harvested the rhizosphere population does not immediately disappear; activities in the vicinity of the roots remain intense though now root decomposition occurs. Any concentration of microorganisms in soil is an indication of the presence of available energy material. The high concentration of organ- isms in the vicinity of roots is not easily explained unless normal plant roots in soil lose soluble organic substances. There are, of course, ab- raded root caps and sloughed-off epidermal cells, but these seem in- adequate to account for the distribution of rhizosphere organisms en- countered around roots of all ages and all species, and for the clear qualitative differences that are detectable among the rhizosphere popu- lations of different species. Root excretions of various sorts have been detected in a number of special cases. Toxic secretions have been said to be responsible for "soil fatigue," and for the effect of one crop on another succeeding it, though microbiological and nutritional factors have not always been excluded (14). Root excretions, either toxic or nontoxic, would be likely to be detected in soil only if they were un- available or only slowly available as energy sources for microorganisms. It is now well established that certain inorganic ions may be lost from plants, particularly as maturity is approached, but the technical difficul- ties of studying under aseptic conditions the excretion of organic sub- stances seem to have discouraged investigation. Claims have been made, and some evidence presented, that sugar and carboxylic acids, such as malic acid, may pass out of roots of certain plants, and amino acids out of legumes (5, 8). Certain enzymes, presumably of plant origin, also have been identified in the vicinity of roots. On microbiological grounds there would certainly appear to be support, admittedly inferential, for the view that some energy source, other than cellular debris, originates in the roots of the growing plant and causes the development of the microbial mantle which ensheathes the roots of plants in soil. EFFECTS OF MICROBIAL ACTIVITY ON SOLUBILITY OF NUTRIENT ELEMENTS Microorganisms may affect the nutrient supply to higher plants in a number of ways. Most reviews and textbooks put emphasis on indirect and inadvertent effects on the solubility of many elements. In addition, it is pointed out, quite correctly, that there are a limited number of 174 Mineral Nutrition of Plants specific inorganic transformations of nutritional importance carried out by autotrophic organisms. Elements such as nitrogen, sulfur, and iron in various forms are involved. These are ordinarily relatively easy to determine; the significance of the transformation and its quantitative aspects can be evaluated. Those indirect effects on solubility of other elements said to be due to products of metabolism of microorganisms are much more difficult to assess. The clay components of the soil and the less weathered rock fragments contain many elements required or utilized by plants in varying amounts. Some of these may occupy posi- tions in the lattice of the clay or silicate; others may be cations retained by the clay or organic matter. At various times investigations have shown that the maintenance of carbon dioxide concentrations in soil may result in an increase in the availability of certain elements. Some of these experiments have been quite artificial, in that lengthy extrac- tions with carbon dioxide-saturated water have been carried out. In others, however, gaseous carbon dioxide has been passed through the soil and changes in availability have been measured by plant uptake. More attention has perhaps been given to phosphate in such studies than to other elements, and it appears to be established that increased solubility of phosphate from rock phosphate or basic calcium phosphate may result, especially in neutral or calcareous soils. The amount of car- bon dioxide evolved in soil by microbiological activity is substantial. Under optimum conditions in soils well supplied with organic matter, carbon dioxide evolution may attain rates as high as ioo pounds per acre per day, though figures of 20-30 pounds per acre per day are more general. Any increase in the hydrogen ion concentration might be expected to increase the solubility of bases such as potassium and mag- nesium, but, as Dr. Jenny indicates, it is not- -now assumed that these must be in the soil solution to be available to the plant. Moreover, ex- periments with plants growing in electrodialyzed colloidal clay have shown a greater availability of certain elements, such as iron, alumin- ium, and silicon as measured by plant uptake, than pass into solution as a result of long carbon dioxide saturation of the clay suspension. Much is sometimes made of the presence of organic acids in soils. Organic acids, however, are not a usual product of bacterial activity under aerobic conditions. They may be produced by certain fungi, but A. G. Norman 175 as they themselves are readily available sources of energy other organ- isms would immediately accomplish their oxidation. Under anaerobic conditions, on the other hand, soluble organic acids may accumulate since they form characteristic products of the activity of facultative and anaerobic forms. Polysaccharides containing carboxyl groups, such as the polyuronide gums produced by some bacteria which utilize cel- lulose, may represent the only "acids" that accumulate under aerobic conditions (10). Continually cropped soils in humid regions become increasingly mom acid. The change in base status in such soils is due not only to cation:, removal in the crop and to the formation of carbonates or bicarbonates; the nitrification process, which is believed to be exclu- sively microbiological, is in part responsible. Ammonia, liberated from organic sources, which at first probably occupies positions on the base exchange complex, may be oxidized to nitric acid and then combine with other bases. The oxidation of 100 pounds of nitrogen, an amount often exceeded per acre per season, would result in the formation of 450 pounds of nitric acid. Similarly, the oxidation of the sulfur present in proteins and thio compounds 'results in the formation of sulfuric acid. It is unlikely, however, that the direct and indirect effects of microbial activity on nutrient availability that result from the production of car- bon dioxide, carboxyl groups, or inorganic acids rival in importance those transformations that are involved in the decomposition of organic matter and plant residues. ORGANIC MATTER DECOMPOSITION AND MICROBIAL SYNTHESIS The soil population is predominantly heterotrophic and accordingly the supply of energy is of major significance in determining its over-all activity. Soil organisms constitute an important link in the chain of carbon transformations in nature. The green plant accomplishes the synthesis of a great diversity of organic substances from carbon dioxide and water; the soil organisms carry out the reverse process and oxidize to carbon dioxide and water almost any carbonaceous material that finds, its way onto or into the soil. That these downgrade processes are accompanied by synthesis is often overlooked, yet the synthetic activi- 176 Mineral Nutrition of Plants ties of the soil organisms are of great significance in determining the availability or supply of certain plant nutrients. The oxidation of any organic compound resolves itself into two downgrade phases, first, the oxidation of the material itself and, second, the oxidation of the mi- crobial cell substance synthesized by the organism. The chemistry of soil organic matter may, therefore, relate more closely to the chemistry of microorganisms than to that of plant residues. If highly available energy material is added to soil, the primary oxidative phase is very rapid with concurrent synthesis of microbial cells; the second phase involving the decomposition of the cell material is slower. Because nitrogen is required for protein synthesis, the nitrogen transformations that accompany synthesis and decomposition can be used within limits as a rough measure of the cell substance present. The amount of cell tissue synthesized per unit of carbon oxidized is not constant with all types of organism. In general, fungi assimilate and utilize for structural purposes a higher percentage of the substrate carbon than do bacteria, and this group has in consequence a higher demand for nitrogen for protein synthesis. When relatively mature but reasonably available plant residues are decomposed by a mixed flora containing fungi, actinomycetes, and bacteria, there appears to be a demand for about 1 .2-1.3 Per cent °f nitrogen to supply the protein needs of the population that develops. If the nitrogen content of the plant material is less than this figure, the decomposition rate will be reduced because the shortage of nitrogen restricts the size of the mi- crobial population that can develop. In this case additional nitrogen, if supplied in the inorganic form, will be immobilized and converted to microbial protein to the extent of the deficit. On the other hand, if the plant residues have a nitrogen content in excess of the requirement, nitrogen will soon be liberated in the form of ammonia as the amino acids in the proteins are oxidized. The nitrogen requirements in organic matter transformations have often been described in terms of carbon- nitrogen ratios. Plant residues with a carbon-nitrogen ratio initially in the neighborhood of 35:1 contain adequate nitrogen for decomposi- tion. As the available carbohydrates are utilized, the population that can be supported becomes smaller with the result that some of the nitrogen initially incorporated into microbial tissue is no longer required for A. G. Norman 177 synthesis of new microbial protein and therefore ammonia begins to appear. As a rough generalization, this may take place when the car- bon-nitrogen ratio of the decomposing residues has narrowed to 20:1. In the decomposition of the more mature low-nitrogen crop residues it can be assumed that most, if not all, of the plant nitrogen is con- verted to microbial nitrogen. Only when the nitrogen content of the material is initially in excess of 1.5 per cent is there likely to be apprecia- ble release of ammonia by deamination and then only after several weeks. For immediate ammonia liberation, the initial nitrogen content must be in the neighborhood of 2.5 per cent or the carbon-nitrogen ratio about 18:1 (//). This is usually only realized by young green plants or leguminous residues. Although nitrogen is quantitatively the most important plant nu- trient the supply of which in the soil is substantially affected by mi- crobial synthesis, other elements also are temporarily immobilized in an organic or inorganic form in the tissues of soil organisms and only become again available to plants on the decomposition of these tissues. Inorganic phosphates may be incorporated in nucleic acids and phos- pholipids, sulfates in sulfur-containing amino acids and sulfonic esters, minor elements such as boron and manganese may be retained in com- binations not at present known. Eventually it may be possible to express quantitatively the requirements of microorganisms for these elements in some such general terms as now can be done for nitrogen. The essential point to be made, however, is that nutrient elements which are required by soil organisms are temporarily immobilized through synthesis, and that this fact must not be overlooked in considering the availability and supply of these elements to the crop. In the case of some elements and in some soils, the amounts so immobilized may be small relative to the supply available to the crop from alternate sources. In other circumstances microbial immobilization may represent an important step in the transformations affecting availability. THE NITROGEN CYCLE The nitrogen cycle in soil is almost exclusively a biological cycle. It is not established that there are transformations of importance ac- complished by purely chemical means. There is no mineral reserve of 178 Mineral Nutrition of Plants nitrogen comparable with that of some other elements, though small amounts of nitrate and ammonia may be added in rainfall. The nitrate supply to the plants depends entirely on the pattern of release of ammonia from organic nitrogen in the soil and, as indicated above, this ordinarily means the release of ammonia from microbial tissues or microbial residues. The nitrification process proper — that is, the oxidation of ammonia through nitrite to nitrate — is entirely secondary in importance. Without ammonia liberation there can be no nitrate. A major unsolved problem in soil microbiology and biochemistry is why the release of ammonia from the organic nitrogen complexes of microbial origin is so slow. The soil is indeed a "frugal custodian" of nitrogen. It is fortunate for the farmer that this is the case, so that in any one season only a small fraction of the nitrogen reserves becomes available to support the growth of the crop. An acre of prairie soil may contain many thousands of pounds of organic nitrogen, yet the amount becoming available in a single season may be only 150-200 pounds. Various theories have been advanced to account for the relative un- availability of the soil nitrogen. Interactions with lignin residues or with clay colloids have been suggested, yet none seems entirely satis- factory. In humid regions where leaching occurs, the rate of release — or mineralization — of soil nitrogen may be an important factor in de- termining crop yields. If at all times there is sufficient nitrogen to meet crop requirements, the soil may be judged well suited for any particular crop. If the pattern of release does not fit the particular nitrogen re- quirements of the crop for optimum growth, supplementary fertilizer nitrogen may have to be supplied even though the total amount re- leased during the season may well be above the needs of that crop. This may not infrequently be the case with early planted or quickly grow- ing crops. For example, it is not unusual in the corn belt to see nitrogen deficiency symptoms in small grain in the spring or in young corn in a cool, wet period. Such conditions arise because microbial activities are temporarily limited by the low temperature or inadequate aeration. Rotational systems involving the return of crop residues may con- siderably alter the pattern of nitrogen mineralization and the amount of nitrogen released during subsequent seasons. More attention has A. G. Norman 179 perhaps been given to practices that may supplement the supply of soil nitrogen than to other less obvious steps that, through the agency of the soil population, may affect the seasonal distribution of the nitro- gen released. For example, effects on the succeeding crop caused by the incorporation of crop residues may differ according to the time of plowing. THE PHOSPHORUS AND SULFUR CYCLES Organic phosphorus transformations in soil, previously largely ignored, are now recognized as being of real significance in relation to the supply of phosphorus to the crop. In some respects the soil phosphorus cycle is far more complicated than the nitrogen cycle (12). There is in most soils a substantial reserve of inorganic phosphorus which may be present in one or more of a number of different forms that vary considerably in availability or potential availability. There are, moreover, inorganic reactions that result in the fixation or con- version of soluble phosphates to unavailable forms. The inorganic chemistry of phosphorus and phosphorus compounds in soil is, there- fore, of great complexity. Superimposed on this there are the synthetic activities of plants and microorganisms that result in the immobiliza- tion of phosphorus in both inorganic and organic combinations. The soluble phosphate utilized by the crop may, therefore, originate in less soluble inorganic sources, or be derived from organic plant or microbial sources if the energy status of the soil is such that mineralization can occur. Because much of the organic phosphorus in microbial tissues is in the form of nucleic acids and phosphoproteins, the immobilization of phosphorus in this form is governed by similar considerations as apply in the case of nitrogen, and there is in fact a quantitative rela- tionship between the nitrogen and phosphorus requirements of micro- organisms for protein synthesis. In some plant tissues and notably in cereal crop residues organic phosphorus occurs also in phytin, a com- pound not believed to be synthesized by microorganisms. Like organic nitrogen, organic phosphorus in soil is less available than might be ex- pected, having in mind the chemical nature of the groupings con- cerned. Phytin appears to be precipitated in acid or alkaline soils as insoluble phytates of low availability (/), though phytin itself when 180 Mineral Nutrition of Plants added to soil may be readily and rapidly decomposed. Nucleotides or nucleic acid fragments may enter into some form of combination with the clay colloids so that availability is much reduced. It has not as yet been found possible to indicate the relationships between carbon and phosphorus that determine whether mineralization or immobilization of phosphorus may occur during the decomposition of crop residues. The phosphorus content of many plant materials is in excess of microbial requirements, but, because inorganic phosphate is normally present in both, no clear picture of the transformations which take place can be obtained. Indeed it is conceivable that the mineraliza- tion of nucleic acid phosphorus is controlled primarily by the nitrogen status of the system, and that this phase of the phosphorus cycle may often be subservient to the nitrogen cycle. Although sulfur is quantitatively far less important nutritionally than phosphorus, there are certain similarities between the phosphorus and sulfur cycles in soil. Like phosphorus, sulfur is involved in organic combination in the proteins of plant and microbial tissues. It also occurs in a variety of thio and sulfonic compounds. Again like phos- phorus, sulfur is present in soil also in various inorganic forms, not all immediately available. Sulfur differs from phosphorus in that through the agency of certain autotrophic organisms it may undergo a series of oxidative reactions which terminate in sulfate. In the case of this element, as with nitrogen, the fully oxidized form is available to the plant so that the normal oxidative biological reactions that occur in soil are advantageous in so far as the availability of these elements is concerned. That this is not always the case will be seen later. OTHER MAJOR AND MINOR NUTRIENT ELEMENTS Much more is known of the role of soil microorganisms in the nitro- gen and phosphorus cycles than in those of some other elements es- sential for plant growth. Because of the relatively high requirements of crops for these two elements, the transformations accomplished by microorganisms may more frequently determine the available supply and, therefore, the total growth than in the case of elements which do not enter into organic combinations or which are required in lesser amounts. A.G.Norman 181 The problem in the case of elements such as potassium or calcium which have cationic properties and fixed valences is relatively simple. Such elements presumably undergo immobilization by microorganisms when they are present in living tissues and then are not freely available to plants. On death of the tissues they can be immediately reutilized. Even so, there are circumstances in which the supply of such elements may be affected by practices which involve microbiological activity. On certain plots at the Long Ashton Research Station in England, apple trees under clean cultivation develop acute potash deficiency symptoms which do not occur if a grass cover is maintained. Other cationic elements of variable valency may similarly be in- corporated into microbial tissues but in addition may undergo biologi- cal oxidations or reductions that may completely change the availa- bility of the element. Manganese is a particularly good example of this group of elements. Exchangeable manganese may be oxidized by heterotrophic organisms to manganic oxide and manganese dioxide (9), both of which are presumably unavailable to plants. Little is known of the microbiology of the oxidative process and still less of the reducing conditions that might make oxidized manganese again avail- able. There is some evidence that temporary reduction can be effected by the addition of thiosulfates, which concurrently are biologically oxidized to polythionates and eventually to sulfate (13). It is to be noted that this class of transformations is not related quantitatively to the needs of the microbial population in the synthesis of cellular material, as is the case with nitrogen and phosphorus. Iron also may undergo oxidative or reductive transformations in soil. In this case autotrophic organisms are known which are capable of effecting specifically one oxidative step but iron oxidation is probably not limited to autotrophs. No doubt it will be found that other micronutrients such as copper, zinc, and molybdenum, whether or not essential to the growth of microorganisms, similarly undergo transformations that are brought about by microorganisms, and which have effects on availability. In- deed, there is already some evidence of this nature in the case of zinc (j). Nothing is known as to the nature of transformations that may be undergone in soil by ampholyte elements such as boron, aluminum, 182 Mineral Nutrition of Plants and silicon. The available boron has been shown to be correlated with the organic matter content of soils but this is hardly sufficient evidence on which to base a claim that the boron cycle must include steps brought about by microorganisms. SUMMARY In attempting to summarize the relationships between the micro- organisms of the soil and the availability of the nutrient elements, major and minor, required by the plant, one immediately encounters the difficulty that there is no simple consistent pattern and, therefore, the relationships cannot be neatly defined in a single polysyllable. The soil population is extremely complex. Its activities are not directed towards meeting the needs of the plant. The tempo of its activities is in part controlled by some of the physical factors that determine plant growth but the dominant factor is the energy supply. Although certain changes in the solubility of some plant nutrients may be the resultant of the production of carbon dioxide and other acids, direct effects are of much greater importance. Many plant nutrient elements are required for microbial growth. All oxidation of organic matter is accompanied by the synthesis of microbial tissues with a concurrent demand for essential elements which are utilized and retained, some in organic combination and some only in inorganic forms. The subsequent re- lease of these elements, particularly those such as nitrogen, phosphorus, and sulfur from proteins and nucleic acids, may largely determine the status of the soil with respect to the availability of these elements. Cationic elements having variable valences, though not involved in organic combination and not required in more than trace amounts, may undergo oxidative or reductive transformation by heterotrophic or autotrophic organisms with the result that the availability of the element to plants may be greatly changed. The role of microorganisms in relation to micronutrient supply in soils has not yet been given sufficient consideration. REFERENCES Since this paper is of a general review character, no attempt has been made to support it by citing a comprehensive list of references. Those that follow may A. G. Norman 183 form a basis for further reading or relate to recent or not well-recognized infor- mation. 1. Bower, C. A., Soil Sci., 59: 277 (1945). 2. Clark, F. M., Adv. in Agronomy, I: 241 (1949). 3. Hoagland, D. R., Chandler, W. H., and Stout, P. R., Proc. Am. Soc. Hort. Sci., 34: 210 (1936). 4. Katznelson, H., Lochhead, A. G., and Timonin, M. I., Botan. Rev., M:543 (i948)- 5. Loehwing, W. F., Botan. Rev., 3: 195 (1937)- 6. Lyon, T. L., and Wilson, J. K., Cornell Agr. Expt. Sta. Mem. 40 (1921). 7. McCalla, T. M., /. Bact., 40: 33 (1940). 8. , /. Bact., 40: 23 (1940). 9. Mann, P. J. G., and Quastel, }. H., Nature, 158: 154 (1946). 10. Norman, A. G., and Bartholemew, W. V., Soil Sci., 56: 143 (1943). 11. Parberv, N. H., and Swaby, R. J., Agr. Gaz. N.S.W., 53: 357 (1942). 12. Pierre, W. H., /. Am. Soc. Agron., 40: 1 (1948). 13. Quastel, J. H., Hewitt, E. J., and Nicholas, D. J. D., /. Agr. Sci., 38: 315 (1948). 14. Shull, C. A., Plant Physiol., 7: 339 (1932). 15. Virtanen, A. J., v. Hausen, S., and Karstrom, H., Biochem. Z., 258: 106 (1933). 16. West, P. M., and Lochhead, A. G., Soil Sci., 50: 409 (1940). MECHANISM OF ENTRY AND TRANSLOCATION OF MINERAL NUTRIENTS IN PLANTS CHAPTER O The Nature of the Process of Inorganic Solute Ac- cumulation in Roots T. C. BROYER T he economic aspect of this symposium, sponsored by the University of Wisconsin, widely known for its agricultural research, and by the National Fertilizer Association, is the seeking of information on practices toward adequate production of plant material, for food, shel- ter, etc., most reasonably under favorable conditions. The mineral nutri- tion of plants is one phase of a larger field of interest, recognized by the organizers of the symposium, viz., plant-soil interrelations. That part dealing with the soil as a medium for plant growth, is covered in earlier discussions. Similarly, certain parts of the mineral nutrition of plants — for example, essential elements — are presented elsewhere. In addition to information on physico-chemical relations in soil and on the quali- tative and quantitative supply of materials therefrom, knowledge of the various aspects of the role which roots play in crop production is im- portant. In particular, therefore, this discussion will present some infor- mation now available on the nature of the process of inorganic solute accumulation as well as pose problems for future investigation.* *A detailed complete survey of the field is impossible in the limited space of this treatise. Here, a number of important aspects are discussed using illustrative material less frequently summarized [compare Hoagland (28)]. The work of several investigators is infrequently and only briefly quoted, since it is anticipated that other authors in this monograph will discuss directly or indirectly in more detail from their own studies and those of their colleagues (e.g., Lundegardh). Most data presented herein are restricted to observations on roots. However, many important researches on absorption have been performed with other tissues. For these the reader is referred, for example, to the literature by Arisz, Brooks, Oster- hout, Robertson, Steward, Stiles, and their associates. 188 Mineral Nutrition of Plants HISTORICAL OBSERVATIONS, DEFINITIONS Probably the earliest experimentation on this subject was by Wood- ward (94) in 1699. He observed the growth of plants in various types of solution, concluding that "earth, and not water, is the matter that constitutes vegetables." Later, investigators recognized the limitation of this conclusion and each added to the realization that "plants are made of chemical elements obtained from three sources: air, water, and soil; and that the plants grow and increase in size and weight by combining these elements into various plant substances" (29). Advancement was attained through the efforts of such men as De Saussure (22), Liebig (40), Boussingault (4), and Sachs (7/). These physiologists relied as usual on the progress in knowledge afforded by their contemporary physico-chemists. This historical advance in knowledge is ably outlined by Reed (6y) in A Short History of the Plant Sciences. Absorption and accumulation Numerous early analyses of plants showed the composition of their various parts. Although not clearly recognized then, it might have been inferred from such data that inorganic solutes were accumulated within the plant from external sources. Probably the first clear evidence of accumulation came from observations of Wodehouse (9^). Affirmation of inorganic solute accumulation came rapidly through observations of many workers including Cooper and Blinks (20), Brooks (6), Col- lander (/6), Hoagland (28), Osterhout (5^), and their associates. All these studies, for expedience, were performed with algae from marine, fresh, and brackish water. A summary of some of these data is pre- sented in Table I (see 28, 5^). Permeability Similar, specific experimentation on inorganic solute absorption was concurrently and subsequently pursued with other plant materials. Often the results were included under the title of permeability studies. These earlier experiments included the course of absorption with time, effects of external inorganic solute and hydrogen ion concentration, temperature relations, and the like, with isolated blocks of "storage" T. C. Br oyer 189 tissue (52, 81, 82), epidermal tissue (24, 34), whole (intact) plants (5"?), and roots (38, 66). Such results have been reviewed by Brooks and Brooks (9) and Stiles (So), and in annual reports (86) entitled "Perme- ability" or "Mineral Nutrition of Plants." The condition of permeability in the strict sense is an important one, for, with a surface limiting entry of materials, the broader aspect of inorganic solute and/or water absorption may be controlled or re- stricted. In the earlier literature the term permeability was broadly used. TABLE I Differential Accumulation of Ions in Halicystis, Nitella, and Chara Halicystis Sea Osier- Pond Brackish Ion Water Oralis houtii Water Nitella Water Chara M M M M X io3 M X io3 M X io3 M X io3 CI 0.52 0.55 0.60 0.90 90.8 73 o 225.0 Na 0.50 0.22 0.56 0.22 10.0 60.0 142.0 K 0.01 0.32 0.006 0.05 54.3 1.4 88.0 Note: Comparison is made of the concentrations of the principal ions between the vacuolar sap and the external bathing medium under natural conditions. However, with advance of research, more specific terminology has been applied. Some of these terms have been defined in recent reviews (10, 11) and as such will be generally used in this treatise. It has been recognized for some time that limiting surfaces or mem- branes are involved where studies on absorption are concerned with living cells and tissues. Several theories have been proposed for the dif- ferential permeability of living surfaces relative to penetration of solutes. Studies on the entry of dyes and relatively nonpolar compounds led to the lipoidal theory propounded by Overton (55). This is a solution hypothesis of permeability, assuming as it does that the penetration of different substances runs parallel with their solubility in lipoidal sub- stances of which the limiting surfaces are supposed to be composed. However, there are compounds which are soluble in lipoidal substances, and yet have been found not to enter living cells with significant rapid- 190 Mineral Nutrition of Plants ity. The solution theory which has been modified in numerous ways, is now combined with another concept, that of ultrafiltration through a sievelike nature of the surface. In the latter hypothesis, the rate of migration of a substance into a cell is related entirely to the size of its existing particles or aggregates, the limiting surface serving as an ultra- filter. Early studies of Ruhland (jo) suggested a close correlation be- tween the rate of movement of nonelectrolytes into cells and their molecular diameter. However, this relation was later reinvestigated by Collander and Barlund (ig). Little correlation was found between per- meability and molecular size. Lipoid solubility is recognized as a prime factor by Collander; however, he considers it incomplete in so far as it merely indicates what substances will enter cells easily under certain conditions. Although probably important, at least in some cases, ultra- filtration likewise cannot be accepted as a complete theory of cell per- meability. Making use of the colloidal character of protoplasm, Clowes (75) has proposed an effect of electrolytes through phase inversion, on cell per- meability. Here, it is conceivable that since protoplasm can be regarded as an emulsoid colloidal system, there will be a continuous phase of one composition (dispersion medium) through which is dispersed at least one other phase, the particles of which are discontinuous. It is supposed that changes in permeability are brought about by the continuous and discontinuous phases changing places, that is, by phase inversion. Penetration would take place only through the continuous phase. The hypothesis would be equally applicable to a solution or a chemical combination theory of permeability. This hypothesis in part may be considered suggestive of others included under theories of chemical combination or adsorption. Some investigators ($8, ^], 83) propose that the substance combines chemically with, or is adsorbed on, a constituent of the protoplasm at its limiting external surface. Absorption would be effected through disturbance of the equilibrium between different pro- toplasmic constituents, with the result that the chemical or adsorption union is subsequently broken again and the substance is released at other places within, or from other parts of, the cytoplasm. Extension and modification of these concepts have been made by Briggs (5), Brooks (7, 8), and others (2, 42, 43, 45. 68). Some of these investigators T. C. Broyer 191 visualize a mosaic structure of the protoplasmic surface in space, while Briggs has proposed a structural difference with time (here especially for cation vs. anion absorption). Phase inversion or time change in struc- ture would suggest that permeability changes might be sudden and the delicate balance would in all probability be very easily disturbed so that the whole of the protoplasm would go to one or the other of the con- ditions. Such a weakness is not inherent in the mosaic structures in space, although the latter may be difficult to envision. The viscosity and electrical charge of the limiting surfaces have been considered by some to be of importance in permeability relations (74, 42). Brooks and Brooks (9, page 138) quote researches indicating that the rate of pene- tration of nonelectrolytes into cells is more rapid than that of electro- lytes; however, little direct comparison has been made. That perme- ability to electrolytes relative to the rate of growth is significant and appreciable is evident from the inorganic composition of plant tissues. Any general theory must account for the entry of ions or ion pairs as well as that of acid and basic dyes or compounds of low dissociation in water. Certainly no single permeability hypothesis thus far proposed will account for the penetration of solutes into living cells. Probably several, if not all, of the suggested theories are simultaneously involved in varying degree, depending upon the circumstances within the spe- cific protoplasm concerned and the substance under absorption study. Some detailed reviews have been made by Brooks and Brooks (9), Stiles (80), and others (jy, 33, 5J, 54, 86). ROOT ANATOMY Relatively little research has been made on the relationship between root anatomy and the physiology of absorption. Some early anatomical work was done by Rufz de Lavison (69). His observations caused him to consider the endodermis of importance as a limiting surface for the penetration of solutes to the vascular tissues. Several studies were made by Priestley and his associates on the physiological anatomy of roots. Priestley and Tupper-Carey (65) suggested the relative impermeability of the growing point of the root to water and substances in aqueous solution. In other publications (64, 72), this point was further stressed as well as placing restriction to movement across the root on the dif- 192 Mineral Nutrition of Plants ferentiation of the endodermis, especially the development of the suber- ized Casparian strip. A transverse section of a typical root through the transition region of elongation and maturation is reproduced from Priestley (6j) as Figure 1. Certain features may be particularly ob- served, namely, the relative areas of cortex and stele, the lack of inter- cellular spaces between cells and vessels within the stele and their prom- inence in the cortex, the differentiation of the endodermis and passage cells therein, and the nature of Casparian strip development and the development of root hairs which increase in number in the region of maturation. These anatomical features no doubt play a part in deter- mining the environmental conditions and the rates of processes within the various tissues of roots. Longitudinally, root structure is distinguished by three portions which grade from one into the other in passing from the tip to the root-stem junction; the extent of each depends upon the relative rates of pro- cesses within and differentiation of the component cells and tissues. These three, respectively, have been termed regions of meristematic development, elongation, and maturation. The longitudinal gradation of effectiveness for penetration and movement of materials across tis- tues is reviewed by Wiersum (92). Movement of solutes across the meristematic region and along its walls is possibly restricted by the close packing of cells, the walls of which are reported to be impreg- nated with protein. Most effective independent flux is considered to occur into the region of elongation. Attended by progressive differenti- ation in tissues farther from the tip, the rate of transverse independent migration of solutes to the stele is considered to be reduced again. How- ever, in this region of maturation, rates of solute influx to the stele may be modified and increased where root laterals are initiated from within the pericycle, thus effectively "piercing" a possible endodermal barrier. The same general pattern has been pictured by Popesco (60). His studies have laid more particular stress on the development of root hairs in the region of maturation. Many investigators have questioned the importance of root hairs in absorption since their pronounced effective- ness is limited to conditions in soils where they may effectively increase the external surface to media of low moisture and/or deficient in- organic solute content. *b \—a 1 C J*y r fejBg^ Ufe. *£t> ^aC^ \ l 1 % i 'i' ?5T 1 1 >; ••: [■; 5 >: £-; e ■$.■■{ * \j;KJj ^2 u s ■:. .-. £ 5 u' :-:-:- :■:?::■:; -HS' « U ' ' rijS;; £>£ B * jfiLi; |:;ig f «» Jj i ;: £ i? j * •:-•-■ *-.. e 1 * SJ- llili ••. n Si;;;: i •J»:':i i >0§ Figure i. Diagram of root structure and endodermis. Transverse section showing (a) endodermis and (b) Casparian strip; (c) median longitudinal section; (d) endodermis with continuous Casparian strip round trans- verse and radial longitudinal walls. — Priestley (63) 194 Mineral Nutrition of Plants A diagram of the spatial relations of different regions in a root tip of tobacco is reproduced from Esau {23) as Figure 2. In this figure, the endodermis is indicated only at the levels where Casparian strips are present. The endodermis, as a cell layer, however, becomes defined be- fore the sieve tubes differentiate. This is clearly shown in the upper of two photomicrographs of a root tip by the same author and reproduced here, as Figure 3 (possibly unusual to a certain extent due to a virus in- fection). Anatomical features of roots relative to migration and accumulation of inorganic solute have been discussed in detail by Prevot and Steward (6/). Their studies on distribution of accumulated inorganic solute will be referred to later, but their anatomical diagram is reproduced at this point (Figure 4). The importance of the cortical tissue is emphasized. MODES OF INORGANIC SOLUTE INFLUX Several modes of inorganic* solute entry into cells may be distin- guished. These are of passive and active nature depending upon whether metabolism of the organism is involved directly or indirectly in the migration (//, 59, j6, 92). Four principal means of movement are at present recognized. These include simple and Donnan diffusion, exchange adsorption, and active or metabolic accumulation. These have been schematically represented in Figures 5 and 6 (//, 92). Detailed discussions of these mechanisms have been published elsewhere (2, 11, 42,53,54,76,92). Diffusion and Donnan equilibria The most obvious mode of migration is that by diffusion. Where- ever a difference of concentration (or better, activity) of a solute exists between two regions, there will be a tendency for movement toward the lower values approaching equality of concentration. This mode of flux across the differentially permeable cytoplasm may be very slow, at least with cells of certain species. The net flux of sodium into Halicystis ovalis is a case in point (Table I). Obviously, for inorganic solutes, a continuous aqueous medium is essential. It was recognized *The movement of organic ions through protoplasm interposed between two solutions needs detailed study. protcxj//e/n po/es protop/?/oe/n po/es er/cj/c/e mature xj//e/r? e/e/ner?£ cortex ep/dier/n/s - e/?oJoafer/7?/s /mmoti/re xg/e/77 e/ement w/t/i secondary tva// SSCUl mature s/ere /i/6e /m/nature s/ere tude rootcap Figure 2. Diagram of a tobacco root tip, showing spatial relations of different regions of the root and of the first vascular elements. This illustrates the order of maturation of tissues in the root tip. In this example the xylem matured at more than twice the distance of the sieve tube from the initials. The en- dodermis, as a cell layer, becomes defined before the sieve tubes differentiate (see Figure 3A), but develops Casparian strips slightly below the level where the first xylem elements mature. In this figure the endodermis is indicated only at the levels where Casparian strips are present. — Esau (23) 196 Mineral Nutrition of Plants Figure 3. Longitudinal sections of the apex of a root of curley-top to- bacco. Details: c, regions of cortical initials; e, endodermis; p, pericycle; s, region of stele initials. A, above, shows spatial order of differentiation of endodermis and phloem relative to the meristem. B, below, shows early differentiation of cortex and stele from meristem initials. — Esau (23) 2 5 315 E o (M 38 E o M 20 495 E o Figure 4. Cellular differentiation in median cross sections through 3 parts of barley roots and the relative accumulation of bromide in them. Right: Anatomy of barley root ("low- salt" type), at different distances from apex. Nos. 1-3, cortex and piliferous layer with- out suberized exodermis. No. 4, 0.6 mm. from apex — differentiation of vascular strands quite evident. No. 5, 5 mm. from root apex — unsuberized endodermis. No. 6, 1-3 cm. from root apex — endodermis suberized but with passage cells opposite xylem groups. No. 7, 3- 5 cm. from root apex — endodermis suberized, without passage cells, and, at extremity of segment, also thickened. Up, piliferous layer: E, endodermis; x, xylem; Ph, phloem; P, pas- sage cells; C, central cavity (axile vessel). Left: Absorption of bromide by segments, ex- pressed in m./e. per 1000 gm. fresh weight. External medium, 5 m./e. per L. of KBr. "Low-salt" roots (white area) from plants 9 days old, grown in distilled water; absorp- tion period 24 hours. "High-salt" roots (blackened area) from plants 15 days old, grown in nutrient solution; absorption period 23 hours. — Prevot and Steward (61 ) i98 Mineral Nutrition of Plants < uj a 5 D a UJ 5 \ r I / _j < z or T UJ V 2 \ //^ Ai/V^ -—-#-4 i"-f UJ \ Z41 \ *~Q 5 r h z Ld 5 (.0 < a O A UJ H > > o (J :• UJ h- D J O (X) . _ -TJ H -? — ul u. o l/l UJ o O J _J 5 < UJ CD [/: O u, O Q — »\< , \uj a -- j_ 1 t\\< 1 rVVj > D Cj u; Q — » \ < _ , \ UJ < Z J 1 \\\ u. H X u_ c 3 < UJ c 0 S3 Q c 03 Oh O -^ _c — C CJ E hj > o 1> V u -G 03 'iX O u 4>J o O bJQ s 4-> ^G 03 o <3J G l-i ■,_l »** iT *^ u <-> c , a > c <-> OJ '"o 'a G en . 03 « be c G X oj 3 > cg -a G o3 —, -o G O O w -O u_, o3 -4—1 G rt 03 i_i TD .5 o3 4-1 03 _r _G 03 x 75 C T3 -G ^ 3< G G _o *4-< (J 03 l-H cortex endo- ' dermis stele wood- tissue bJ. i = I b -r 1 eel with vacuole „.JT. V JL .. ''A ■ '///'/<. yy-A ■■■ H if -y" ; 'v < ■■■//■y/'///4 m * ■ ■ / ■■;■'■ W6 =fr polar.active transfer > passive movement. diffusion Figure 6. Scheme illustrating the theoretical considerations con- cerning lateral transport through living root tissues.— Wiersum (92) 200 Mineral Nutrition of Plants early by Donnan that under certain circumstances a modified equili- brium of diffusion may obtain. This phenomenon occurs where one of the ion pairs of a compound in solution does not readily penetrate a differentially permeable membrane interposed between two aqueous solutions. The equilibrium conditions imposed by this phenomenon are such that, for monovalent ions for example, the product of the molar concentrations of the cation (C+) and anion (A~ ) pairs in the one medium equal the same product in the solution on the opposing side of the limiting surface, i.e., (C+)i X (A~)i = (C+)e X (^_)e- Ex- pressed otherwise, at equilibrium the ratio of the concentrations of the diffusible cations in the two media should equal the inverse ratio of the diffusible anions, i.e., (C+)j/(C+)e = (A~)v/(A~)i. Studies in vitro have shown this phenomenon to be qualitatively valid; however, in nature the ratios required at equilibrium are not usually observable. The latter observations indicate that although generally applicable, ionic equilibria of this sort are not universally attained in systems with living cells. Compare the relevant inverse ratios of cation to cation and anion to anion, particularly for K+, Na+, and Cl~, for Halicystis, Nitella, and Chara in Table I. One must infer, therefore, that the Donnan equilibrium as originally proposed is not to be regarded as of primary importance between at least some living cells and their natural environ- ment (//, 54). Exchange adsorption With living tissues, we are dealing with colloidal systems, and the phenomenon of exchange adsorption can play a role in cellular physi- ology of inorganic solute absorption. Ions are adsorbed on the surfaces of charged colloids. These may be exchanged for ions with the same sign from the dispersion medium. This mode of ionic migration has been extended recently to include an exchange of ions between colloidal micelles where interpenetrating spheres of ion oscillation exist. Thus, also, the movement of ions along surfaces, without intermediate passage into the dispersion medium, may be effected. These exchange phe- nomena are schematically represented in Figure 7, from Jenny and Overstreet (j6, ^7). Such movements are discussed in detail elsewhere in this book. T. C. Broyer Active or metabolic accumulation 201 Simple diffusion leads toward an equilibrium in which a substance in solution tends to approach equality of concentration within the system. Under appropriate conditions, Donnan phenomena and/or ex- change adsorption may lead toward a condition in which differential accumulation of inorganic solutes may obtain within cells. However, some evidence would suggest that these modes of migration are not all inclusive. A complementary process has been proposed and recognized (//, 2j, 28, 42, 59, j 5, 86), whereby inorganic solutes at least are ac- ;^ © Jolution - exchange Jur/ace m/pmtm • ! 1 1 : O x:l - t- -1 1 lJ "6" • 1 1 • @ ■L-_- — • r- © L)-- Contact exchange '/jf/uf/on m^e/j Figure 7. Diagram of various modes of ionic migration, especially adsorption exchange. — Jenny and Overstreet (jy) cumulated within cells of plants. This mode of movement is considered to be directly related to metabolism of the cell and is a polar or "unidi- rectional" migration. Each and all of these modes of accumulation, induced as well as active, are probably involved in variable degrees in the net movement into cells and tissues of plants. In this regard, roots of higher plants are of particular concern initially, since they are the organs concerned primarily with absorption for these species. FACTORS DETERMINING RATE AND EXTENT OF INORGANIC SOLUTE ABSORPTION Aeration There are two main groups of influences which determine the rate and extent of solute absorption. These factors may be due either to 202 Mineral Nutrition of Plants external or environmental conditions, or to internal controlling con- ditions. Of the external or environmental conditions, aeration, tempera- ture, and the supply of water and mineral salts are important. These conditions may alter and determine the internal physico-chemical con- ditions and the relative rates of processes, some interrelated with absorp- tion, in roots. Studies with abscised roots of barley (jo) and oats (59), and roots of decapitated wheat plants (42) have clearly shown the necessity of an aerobic environment for the accumulation of salt in such organs. Experimental results with abscised roots of tomato and rice, as well as barley, are presented in Figures 8A and 8B, from experiments of Vlamis and Davis (88). Here, the reduced accumulation of potassium and bromide at oxygen values below 3-5 per cent is clear. A fairly close correlation exists between these effects of oxygen supply on ac- cumulation, and on total respiration by comparable tissues. This is evident when the respiration curves shown in Figure 9 are compared with the accumulation curves of Figures 8A and 8B. Similar concord- ance between respiration and ion accumulation rates in potato roots has been reported (77). The relationship between these factors will be dis- cussed further in another section. Temperature The temperature of the environment is a second factor which may control the rate of physico-chemical processes. As with aeration effects, a similar correspondence between the effects of various temperatures on the processes of accumulation and respiration have been obtained. Several investigators have reported these phenomena with roots (30, 42,84,91). Such effects may be exemplified by data derived from ex- periments of Ulrich (84) with abscised roots (Figure 10). In these histograms, it may be observed that increases in both accumulation and respiration rates are associated with rise in temperature of the medium bathing the tissues. More than doubled increases in rates for a ten- degree rise in temperature are indicative of these processes being of physico-chemical nature. From the concordance of temperature effects obtained between accumulation of inorganic solutes and respiration (especially the enhanced respiration), investigators have generally in- ferred some interrelation between them. This is quite possible, since T. C. Broyer 203 D. < CO Q Ld CO CO UJ i- Or $L o_ — ?J z S F * z Ld u z o a < 0 a iii co CO l_ Ld £) IX -= n L_ X « Ld to ( z f.l n ■> z i 0 ff or 1 I- Z ill u e barley 0—0 fomato 7—9 rice cv control value 8 10 12 14 OXYGEN 7. 18 20 100 100 OXYGEN % Figures 8A and 8B. Effect of aeration on salt accumulation by abscised roots of barley. Plants 4 weeks old; abscised roots in experimental culture solution (KBr, 5 milliequivalents per liter) for 24 hours. Above: potassium absorption; Below: bromide absorption. — Vlamis and Davis (88) 204 7 O h- Q_ 100- *> ~) CO z o 80 u z UJ 6Q o > X o 40- UJ > h- < t>() _J UJ Q: oJ Mineral Nutrition of Plants barley o— © tomato 9— v rice 0 8 10 12 14 Id" 18 20 100 OXYGEN % Figure 9. Effect of aeration on respiration by abscised roots of barley. Conditions of growth and experiment as in Figures 8A and 8B. — Vlamis and Davis (88) a 10 y 15 25 35 TEMP. IN DEG. C. Figure 10. Effect of temperature of the bathing medium on inorganic solute absorption and respiration by abscised barley roots. Plant growth, 5V2 days, dark, in a dilute arti- ficial tap water. Experimental solution 0.005 M KBr -|- 0.0005 M CaS04; time, 8 hours— Ulrich (84) T. C. Broyer 205 the Q10 values for both absorption and respiration are very similar in magnitude for like increases of io° C. These vary from 2 to 3 at lower temperature intervals to about 1.5 at the higher. The respiratory quo- tient does not vary significantly over the temperature range studied; similar 010 values are obtained for both oxygen involved and carbon dioxide evolved. The Q10 values for absorption of bromide are con- sistently above those for potassium over similar temperature ranges, the difference decreasing with increase of temperature. If any significant inward migration of potassium by exchange over that between anions is possible, this should tend to bring the Q10 values for potassium, com- pared with bromide, closer together (see later discussion on "anion" respiration). 330 Figure ir. Effects of temperature and nature of the culture medium on the rate of outward movement of radioactive potassium from abscised bar- ley roots which had previously accumulated radioactive potassium. Plants 4 weeks old. Set A: 0.005/V nonradioactive KC1 at 20.0 ° C; set B: 0.005N nonradioactive KCi at i.o° C; set C: distilled water at i.o° C; set D: dis- tilled water at 20.00 C. 206 Mineral Nutrition of Plants In the immediately preceding discussion, the pronounced effects of the external conditions of aeration and temperature on the over-all process of salt accumulation were considered. Aside from the possible influences of aeration and temperature upon permeability itself, it would be expected that the temperature condition might also modify the rates of diffusion and exchange. Effects of this sort would be mani- fested through increase of kinetic energy of mobile substances in solu- tion. Since the energy intensity for migration of a solute would be proportional to the absolute temperature, this influence, however, would be relatively small. Figure 12 A. Effects of temperature and aerobic nature of the culture medium on the rate of inward movement of radioactive potassium. Barley plants 3 weeks old. Sets A and B: 0.50 C, C02-free air passed through the culture solution; sets C and D: 20.5 ° C, COa-free air passed through the culture solution; sets E and F: 20.5 ° C, CO,-free N2 gas passed through the culture solution. The ex- perimental points indicate radioactivity measurements of K*C1 culture solutions in which excised roots were immersed. The data are plotted as differences between the initial culture value and the values at subsequent periods of time. T. C. Broyer 207 A study was made (/j) of the possible effects of temperature and aeration on exchange adsorption with roots. The results of these studies are summarized in Figures 11, 12A, and 12B. In Figure 11, the rates of outward isotopic diffusion and exchange of potassium at two tempera- tures are shown. Simple diffusion of salt from roots into distilled water was very slow; exchange of cations was rapid. About 10 per cent of the previously absorbed radioactive isotope of potassium was readily ex- changeable (compare with data of Overstreet and Jacobson, referred to later for other cations and for anions). The rates of ionic exchange were not significantly influenced by temperature. In Figure 12, a similar experiment on exchange at two aeration levels is presented. The upper graph (12A) shows the relative accumulation of radioactive potassium 2 3 TIME IN HOURS Figure 12B. Effects of aerobic and anaerobic culture medium environments on the rate of outward movement of radioactive potassium. Following the du- plicate treatments presented under Figure 12 A, individual sets of excised roots were exposed to culture solutions of o.oo^N KC1 at 20. 50 C. as follows: Pretreatment — Absorption of K*C1: A-B, temperature 0.5 ° C, air-treated culture; C-D, temperature 20.5 ° C, air-treated culture; E-F , temperature 20.5 ° C, N2-gas-treated culture. Treatment during efflux study: A, COa-free air passed through culture solu- tion; B, CO,-free N2 gas passed through culture solution; C, C02-free air passed through culture solution; D, COa-free N2 gas passed through culture solution; E, C02-free air passed through culture solution; F, C02-free N2 gas passed through culture solution. 208 Mineral Nutrition of Plants during pretreatments with two conditions of temperature and aeration. The lower graph (12B) shows the experimental measurements of isotopic exchange under aerobic and anaerobic conditions. The ex- change rate was not enhanced by aeration (compare with upper graph). On the contrary, an even lower rate of exchange was obtained with aeration, which is probably related to concurrent reaccumulation by metabolic influx or internal translocation of previously adsorbed ions from the effective loci of exchange. That exchange does occur is clear from these experiments. It should be mentioned, in passing, that while this limited exchange of isotopic cations was proceeding, a pronounced net influx of total potassium occurred where oxidative metabolism was favored (i.e., with aeration and at higher temperatures). These observa- tions suggest two modes of movement of inorganic solute apart from diffusion phenomena, namely, an ionic exchange, more or less inde- pendent of metabolism, and an accumulation mechanism, the latter directly related to concurrent metabolic activity. Water supply A supply of water to the living system is obviously necessary. How- ever, it may be less obvious that the relative supply may alter or de- termine the rate of internal processes, especially the accumulation of inorganic solutes. The influence is indirect. The movement of water into tissues will dilute the concentration of solutes within the system, thus lowering the accumulation level therein. Since the maximum inorganic solute level of tissues is limited by hereditary potentialities and rates of processes, further movement of such solutes inward may occur through dilution within, or by translocation from, the loci of initial absorption and accumulation to other parts of the plant (30). (Compare with the discussion of internal solute concentration.) If, as is considered to be the case by most investigators, the movement of inorganic compounds is in the form of ions or ion pairs, water is obviously essential to this dissociation. Certainly ion exchange phe- nomena are of this sort. The water factor in mineral nutrition is dis- cussed elsewhere in this monograph by Biddulph, Burstrom, and Wad- leigh. T. C. Broyer 209 Inorganic solute supply The supply of mineral salts is axiomatic. Furthermore, the con- centration and relative proportions of such substances may alter or control internal processes. Again, the inorganic solute status may modify the rates of the two processes earlier referred to, namely, accumulation and respiration. The effect to be noted is that of concentration of a single salt in the bathing medium on the rates of respiration and of accumulation of the ions of the compound supplied. With increasing pH 7.0 6.0 5.0 4.0 0.005 0.025M KBr CONCENTRATION Figure 13. Effects of potassium bromide concentration on the metabolic activity of excised barley roots maintained at 25.00 C. for 8 hours. K and Br values given as milli- equivalents per liter of expressed sap. R.Q. = C02/02; Sugar — total sugars expressed as grams glucose per liter of expressed sap. The following are given as milliequivalents per liter of sap: O.A. (total nonvolatile organic acids) and NH3 -f- amide-N (ammonia plus amide nitrogen). pH values were determined on expressed sap. Conditions of growth and experiment were the same as for Figure 10, except that KBr only was supplied in the experimental cul- ture solution. — Ulrich (85) 210 Mineral Nutrition of Plants concentration of a dilute salt solution, the rate of accumulation is like- wise greater. A corresponding, though not necessarily proportional, increase in respiration is generally observed. These effects are graphi- cally represented in Figures 13 and 14 from Ulrich (85) ; see also Figure 15. In Figure 13 it may be observed that the rates of accumulation of the ions of a particular ion pair supplied (here, potassium and bromide) may be rather similarly affected. However, the relative rates of anion and cation accumulation may be distinctly different. Thus, a monoval- ent ion (cation or anion of a pair) usually accumulates more rapidly than a divalent "partner." This may be seen by comparing the absorp- tion of bromide and calcium from calcium bromide or potassium and sulfate from potassium sulfate shown in Table II (84). Further, the rate of absorption of a particular ion is markedly influenced by the nature of the accompanying ion of opposite charge. Thus, bromide *NH3+ AMIDE -N 0 0.0025 O.OI25M CaBr, CONCENTRATION Figure 14. Effects of calcium bromide concentration on the metabolic activity of excised barley roots maintained at 25.00 C. for eight hours. Data expressed in the same way as in Figure 13. Conditions of growth and experiment were the same as for Figure 10, except that CaBr2 only was sup- plied in the experimental culture solution. — Ulrich (85) T. C. Broyer 211 + z . Z i DC Z er O.A. CON Change m. VI v> -a c U o u a. -9 '5 o < z < c — • C u "-■>. 0 "" u ^ o 6 U E o U _ < 0 dc a. - -- H 3 OC r o a e-i « 0 o — «; .r, 0 O < u « „ 0 — U u * + Os + oo + 1 o 1 O to rj sO t» IO Os Os o oo sO Ti- to ro rO rO to O oo sO Os sO oo oo oo IO Os oo 00 Tt- o Os sO sO 00 00 ti- O sO Os o rO sO O o If) o oo so O O OO 29 r.o w,73 Cftt/3 . ,V) J. CQr^ WU DCr03 I ) W s< % ^^ ^^ >r> io ^ lO O "" ro IT) O N O lO o « o O o o o o o o o o r o o o o o o o o oo DC rt W o "^ m o o o o o •7 a) ^^ ^ iri K") O o o o o ^o DC c/; 6u "^«! N o o o o o OU io io O) O o o o o o - zo a! rt UU ^^ to to M O o o o o oo oo oo oo oo oo oo oo oo >5 3 Os > o o Li 1-1 ■S 0 o *T" c • "" d Ui 3 C O U u o LA (S 00 so d w u -a c a o u ^ .s a cd G CQ O nj ■5 U 5^ 3 u "« C 143 O ,0 <-> * 8 vs u 0 -o C v u. *-< a "0 u "(J 1- x 03 s*. "*" O u VI ^ 3 vi « u "- 3 -a > -a - C v; 3 o — o 1H u 0 S * be o? C OC I- s£ u o -ti q ^ s^ «J - s£> rr. c td c 0C ro V) ro PL, B 3 C O o 1_ ■a > 3 3 %n
  • 13 z 2400 X z o 2000 »- u 3 1600 a o a. 0. 1200 N o o 800 400 INITIAL COHC. OF SUGAR a. < o 20 10 < o 3 •A o 100 '50 PERIOD IN HOURS 200 Figure i8. Time-course of respiration (C02 evolved) and sugar tion in abscised barley roots. deple- in recent publications (g, n). Permeability is obviously particularly concerned with phase boundaries. The permeability is related to the specific protoplasmic organization characteristic of the species and altered, within limits imposed by genetic factors, by the environment. Some of the features of surface limitations have been outlined earlier. A lowering of temperature reduces the rate of metabolism, but it will also increase viscosity. Permeability may be more readily separated from metabolic effects on accumulation, by studies in which the external medium is generally more concentrated than that within the tissues, so that simple diffusion could account for the influx without neces- T. C. Br oyer 221 sarily involving other modes of movement. From data such as those presented in Table VI (32), the marked reduction of the inward migra- tion rate for potassium bromide with lowered temperature is clear. The pronounced effect of deficient aeration in reducing the rate of accumulation of inorganic solute from solutions of low concentration TABLE VI Effects of Aeration, Temperature, and Cyanide Treatments on Salt Movement into Excised Roots of Barley Net Respira- Experimental Conditions Absorption tion Rate K Br mg. C02/ m.e./l. sap g-/hr. Experiment A H20, air — 1.6 o 0.29 H20, N2 — 5.1 o 0.16 KBr 5 m.e., air 29.1 22.3 0.32 KBr 5 m.e., N2 — 0.1 3.2 0.12 KBr 5 m.e. + KCN 1 m.e., air ... — 2.5 o o. 19 KBr 50 m.e. + KCN 1 m.e., air . . 4.2 6.8 — Control roots (initial status) 24 o Experiment B KBr 60 m.e., 0.750 C, air 18.5 16.1 KBr 60 m.e., 0.750 C, N2 5.8 9.7 KBr 60 m.e., 18. 50 C, air 58.4 50.0 KBr 60 m.e., 18.50 C, N2 7.2 10.7 Control roots (initial status) ... 22.0 o Note: Experiment A: plants 3-4 weeks old; experimental period 8 hours, temperature about 200 C. Experiment B: plants 3 weeks old; experimental period 12 hours, temperature 18.50 C. was pointed out under discussion of environmental factors and ex- emplified by the data in Figures External solution 3 3 re > oj -g CO OJ cm CO External solution Rb* or Br* in Co per Min. per M fa- re - 1/3 JJ 70 ' CO OJ cm I) "3 3 (J at > b C c 'Z c c L - c a I 'Z o. C > 3 3 ) ) | 4 J L, \D U^ N VO U-> ►- (N M M t\ « O O 3 oj 3 i- -./ — s. ■ rvj rr; rO rr, t^ i/-\ ~ -1- -»• ir IS 0 00 0 «3- CO >- >-. C3 re -a -* 0 •"■ "-1 4-J -W _£ -C DJQ &C n r -0 -a OJ OJ C3 rt u, 1* OJ OJ re re u 0 * * _Q _Q Oh Dh M rl O O O On f VO CO O 1- -t- 1 1 m rl in r^ On "1 (N O w 1- -*■ 00 a E 1 1 Ih «J c_ X 0 0 W O t" 00 00 ro n (S PI 0 0 ■«t- \r> 00 r^ tv> \o 0 0 -1- m 00 0 r^ rs - 0 in ri s u ui 2 _C J3 3 m « O -3 re 3 O u OJ — re Dh re CO CO > Ui -5 3 3 rt O C 3 3 O O ~ > - re — 6 c OJ l- LI ui OJ s 3 re 3 0 Ui O 1-t - £ u. c_ OJ 5 K OJ Ui a 3 O CO CO 4-* u J3 3 n re OJ b u 1 — 1 □ OJ re - C > — OJ > 0 R > '3 — 3 O 3 re > 4-1 Be ^3 'C 3 3 cr u 6 c OJ 3 — £ OJ 1- > O cr 0^ ~D ITl — 3 "re 3 Ut — CO # e 00 3 CO re * -3 Dh 3 JD .3 Ht £ cm ra 0 re CO 3 kw — 3 re 3 0 3 a Ui OJ — CO re 3 Dh 3 0 5e 3 g 3 3 CO OJ ~ re C D re V) CA O 3 O Dh re T) ca 3 OJ X OJ "re re i Ut OJ Oh 0 a, OJ n — - 3 0) C X U E X OJ - £ 0 OJ 0 OJ — £ Ui OJ Q, X £ 0 H ^I H 0 ij — ■— u Dh U OJ ^ OJ Cfl R — re CO OJ UJ CO 0 a £ - — 3 V c _o — OJ 0 - to _0J -6 3 u j_, — M 3 (N 3 f) u OJ 4-1 (J 4-1 b -0 3 Jj 3 re OJ r, OJ 3 3 Ul O E 4; in 26B, in Sr*Cl2 or KI* solutions. Following each immersion the segments were washed in distilled water and counted (curves A and B). The segments were then transferred to 0.005 AT nonradioactive solution of the same salts and the process of suc- cessive immersions repeated (curves C and D). The amount of radicele- ment in each segment is expressed in terms of counts per second.— Over- street and Jacobson ($6) T. C. Broyer 243 50- 40i br Q Z o u UJ CO or u Q. CO H Z D O u MINUTES Figure 26B. These graphs are discussed under Figure 26A. 244 Mineral Nutrition of Plants may be accumulated throughout the length of unbranched roots of succulent species. Concentrations of bromide at various levels were represented in Figure 4. At the time of measurement, a gradient of concentration was evident, decreasing from the meristematic distal end MM FROM ROOT TIP Figure 27A. Graphs here and in Figure 27B show the magni- tude of nonmetabolic absorption of Rb86, P32, Sr85, and I131 (ex- pressed in counts per second per mm.) as a function of distance from root apex. In 27A, apical segments of barley roots 1 cm. long and approximately 0.4 mm. in diameter were immersed in Rb*Cl or KH2P*04; in 27B, in Sr*Cl, or KI* solutions for 30 minutes at o° C, and subsequently sectioned and counted. The dotted line labeled "outside solution" in each case corre- sponds to the activity of a volume of the bathing solution equal to that of 1 mm. of root segment. — Overstreet and Jacobson (56) toward the proximal or root-stem transition part of the root. A similar distribution was shown spectrographically for potassium and rubidium by Steward, Prevot, and Harrison (79). The similarity in shape be- tween a distribution curve for metabolically accumulated rubidium (Figure 24, from 79) and a respiration gradient curve (Figure 25, from 50) on the same types of roots is probably further evidence for the T. C. Broyer 245 close relationship between these two processes (see also 25). The shape of the distribution curve for adsorption exchange along the same types of barley roots may be slightly different. Curves of this sort are repre- sented in Figures 26A, 26B, 27A, and 27B, from Overstreet and Jacob- son (j5> 56). These show the differential adsorption pattern of various anions and cations in short time experimental treatments. The salt distribution pattern in the plant as a whole will depend on many interrelated factors including relationships between shoot and root, organ and organ, tissue and tissue, and among cells of like type; for the plant is an integrated machine, the process in any one part being dependent upon the concurrent and past action and conditions in every o o Q: 5 (r UJ a a z o UJ and Overstreet, R., /. Phys. Chem., 43:1185 (1939). 38. Kahho, H., Biochem. Z., 123:284 (1921). 39. Kramer, P. J., and Wilber, K. M., Science, 110:8 (1949). 40. Liebig, J. von, Die organische Chemie in ihrer Anwendung auf Agri- cultur und Physiologie (Braunschweig, F. Viewig und Sohn, 1830). 41. Lepeschkin, W. W., Am. J. Botany, 19:568 (1932). 42. Lundegardh, H., Ann. Agr. Coll. Sweden, 8:233 ( x940)- 43- , Ar\iv. f. Bot. 32A, No. 12:2 (1945). 44- , Ann. Roy. Agr. Coll. Sweden, 16:372 (1948). 45- > Discuss. Faraday Soc, 3:139 (1948). 46. , Ann. Roy. Agr. Coll. Sweden, 16:339 (l9$)- 47- > Kgl. Svenska Vetens\apsa\ad. Handl., 1:295 (T949)- 48. Luttkus, D., and Botticher, R., Planta, 29:325 (1939). 49. Machlis, L., Am. /. Botany, 31:183 (1944). 50. , Am. f. Botany, 31:281 (1944). 51. Milthorpe, J., and Robertson, R. N., Australian /. Exptl. Biol. Med. Sci., 26:189 (1948). 52. Nathanson, A., fahrb. wiss. Botan. 39:607 (1904). 53. Osterhout, W. }. V., Botan. Rev. 2:283 (1936). 54. , Botan. Rev., 13:194 (1947). 55. Overton, E., Vierteljahrsschr. naturforsch. Ges. Zurich, 44:88 (1899). 56. Overstreet, R., and Jacobson, L., Am. J. Botany, 33:107 (1946). 248 Mineral Nutrition of Plants 57. Pantanelli, E., Jahrb. wiss. Botan., 56:689 (1915). 58. Pauli, W., Sitzbcr. A\ad, Wiss. Wien, Math-naturw. Klasse. Abt. 1, 113:15 (1904). 59. Petrie, A. H. K., Australian /. Exptl. Biol. Med. Scu, 11:25 (1933). 60. Popesco, St., Bui. Agriculturii, 4:1 (1926). 61. Prevot, P., and Steward, F. C, Plant Physiol, 11:509 (1936). 62. Priestley, J. H., New Phytologist, 19: 1 89 (1920). 63. , /. Roy. Hort. Soc, 51:1 (1926). 64. , and North, E. E., New Phytologist, 21:113 (1922). 65. Priestley, J. H., and Tupper-Carey, R. M., New Phytologist, 21:210 (1922). 66. Redfern, G. M., Ann. Botany, 36:167 (1922). 67. Reed, H. S., A Short History of the Plant Sciences, (Waltham, Mass., Chronica Botanica Co., 1942). 68. Robertson, R. N., and Wilkins, M. J., Aust. f. Sci. Res., Bi:i7 (1948). 69. Rufz de Lavison, J. de, Rev. gen. botan., 22:225 (x910)- 70. Ruhland, W., Ber. deut. botan. Ges., 30:139 (1912). 71. Sachs, J. von, Landw. Vers.-Sta., 2:219 (i860); 3:30 (1861). 72. Scott, L. I., and Priestley, J. H., New Phytologist, 27:125 (1928). 73. Skoog, F., Broyer, T. C, and Grossenbacher, K. A., Am. J. Botany, 25749 (1938). 74. Spaeth, R. A., Science, 43:502 (1916). 75. Steward, F. C, Ann. Rev. Biochem., 4:519 (i935)- 76. , Trans. Faraday Soc, 33:1006 (1937). 77. , Berry, W. E., and Broyer, T. C, Ann. Botany, 50:345 (1936). 78. Steward, F. C, and Martin, J. C, Carnegie Inst. Wash. Publ. 475:87 (T937)- , . . 79. Steward, F. C, Prevot, P., and Harrison, J. A., Plant Physiol., 17: 411 (1942). 80. Stiles, W., Permeability. (London, Wheldon and Wesley, Ltd., 1924). 81. , and Jorcensen, I., Ann. Botany, 29:611 (1915). 82. Stiles, W., and Kidd, F., Proc. Roy. Soc. {London), 690:448 (191 9). 83. Szucs, J., Jahrb. wiss. Botan., 52:85 (1912). 84. Ulrich, A., Am. J. Botany, 28:526 (1941). 85. ■ , Am. J. Botany, 29:220 (1942). 86. Various authors, Ann. Rev. Biochem. (1935, rT). 87. Viets, F. G. J., Plant Physiol., 19:466 (1944)- 88. Vlamis, J., and Davis, A. R., Plant Physiol., 19:33 (J944>- 89. Wanner, H., Ber. schweiz. botan. Ges., 58:383 (1948). 90. , Ber. schweiz. botan. Ges., 58:123 (1948). 91. , Vierteljahrsschr. naturforsch. Ges. Zurich.. 93:99 (1948). T. C. Broyer 249 92. Wiersum, L. K., Rec. trav. botan. neerland., 41:1(1 946-47). 93- Wodehouse, R. P., /. Biol. Chem., 29:453 ( 1917). 94. Woodward, J., Phil. Trans. Roy. Soc. London, 21:193 (^99) 95- Zimmerman, P. W., Am. ]. Botany, 17:842 (1930). 96. Zscheile, F. P., Protoplasma, 11:481 (1930). " CHAPTER / The Mechanism of Ion Absorption HANS BURSTROM D uring recent years there have appeared complete theories of the mechanism of the active absorption of ions, which in a rather definite way have brought together the vast experiences con- nected with the subject. It is an easy task to explain an active absorp- tion and accumulation of one kind of ion, and this can be brought about in experiments with models of different kinds. However, the crucial point in every such theory is to explain the simultaneous absorp- tion of anions and cations and their accumulation together against the diffusion gradient. The first of these theories was advanced by Lundegardh (j), and was founded on experiments in progress since 1933; a second theory, a modification and development of the former, has been advanced by Robertson (6). Although the main features of both theories are proba- bly well known to you, I feel it necessary to review them in some detail before entering into a discussion of the present situation of this field of research. Lundegardh's theory is based on four assumptions, which perhaps today may be regarded as generally accepted facts. 1) The absorptions of anions and cations are independent of each other to such an extent that different mechanisms must be responsible for each. 2) The absorp- tion of cations takes place in two steps, one involving the absorption of ions from the external solution and the other the excretion of the ions from the cytoplasm into the vacuole. The latter step constitutes the real accumulation. A number of recent investigations have cor- roborated this view for different plant materials such as roots, storage tissues, and even leaves. Arisz (/) has shown that leaves of Vallisneria 252 Mineral Nutrition of Plants carry out active accumulation of ions in essentially the same manner as other organs. It has also been verified by several investigators in experiments with tagged cations that the first step of the absorption is reversible and is probably, as was assumed long ago, an adsorption or a tendency toward a Donnan equilibrium on the surface of the plasma colloids. 3) The absorption of anions is of a different nature. It is irreversible and, as has been shown with tagged ions, it takes place not only against the diffusion gradient but also against the charge of the cell (which is predominantly negative) and the adsorption potential. It depends on a portion of the total respiration which is usually called anion respiration and which Robertson has called salt respiration. 4) This anion respiration is different from the rest of the respiration which is called the ground respiration. Some of these points, especially point three, the connection of ac- cumulation of anions with respiration, have been questioned, but a number of new data, particularly from Robertson, support the as- sumptions made by Lundegardh. New evidence in favor of the opinion that respiration is connected specifically with the intake of anions is found in the determinations of the effect of temperature on ion absorption. I may cite the experiments by Wanner (8). These and other investigations without exception show a significantly higher, sometimes very much higher, temperature coefficient for the absorption of anions than for the absorption of ca- tions. Wanner found a 0U) of 1.4 for the absorption of potassium from potassium nitrate but a QU) of about 2 or 2.5 for the absorption of the nitrate ion. With ammonium chloride the corresponding figures were 0.9 for the cation and 1.5 for the anion. Still greater differences have been recorded by Jacobson and Overstreet. A low temperature coeffi- cient is characteristic of physical reactions, whereas chemical trans- formations normally have quotients of two or three. These observa- tions, therefore, support the view that the first step in the absorption is a physical phenomenon, while the second step is connected with metabolism, and that this applies especially to the absorption of anions. Robertson and his collaborators further have verified the point that respiration is directly connected with the process of ion absorption. Not only does the total respiratory rate increase to a higher level if Hans Burstrom 253 salt is added to the external solution, as demonstrated by Lundegardh long ago, but also, if salt is withdrawn from the external medium, the rate again goes down to approximately the initial level. Thus, this effect of salt on the respiration is reversible, which it would have to be, if it is connected with the absorption of ions and not only with the presence of ions in high concentrations within the tissues. This salt respiration has been further defined by its high sensitivity to the heavy metal inhibitors — hydrocyanic acid, carbon monoxide, and sodium azide — to which agents the ground respiration is almost in- sensitive. This point, initially brought out by Lundegardh, has been verified by several investigators who have shown that the intake of ions can be wholly inhibited by hydrocyanic acid. Especially convincing are the experiments of Robertson, according to which an addition of hy- drocyanic acid stops both salt intake and the associated increase in rate of respiration at once, but leaves the ground respiration almost unaffected. Consequently, with roots or tissues in water, hydrocyanic acid has little effect on the respiration. According to prevailing concepts, this would indicate that the salt respiration is catalyzed by hemin compounds, while the ground respira- tion is catalyzed by a different enzyme system, the nature of which is unimportant in this connection. The hemin compounds contain iron as the active constituent with which the mentioned inhibitors form complexes and thus inactivate the enzyme. On the basis of the above evidence, Lundegardh assumed, as did Robertson later, that the salt respiration is catalyzed by cytochrome systems, because cytochromes are probably universally present and contain hemin compounds which are easily inactivated by hydrocyanic acid. Lundegardh and Robertson go even further and claim that the cytochromes themselves play a fundamental role in the absorption of the anions. This remains to be proved, but is at present an integral part of the theories advanced by both men. The active part of the cytochromes is the iron atoms which change their valence by taking up and giving off electrons according to the formula, Fe+++ -f- e~ ' Fe++. Thus the cytochromes serve as carriers of electrons and function at one end of the respiratory system. Now it is assumed, or is really an established fact, that when the Fe+++ 254 Mineral Nutrition of Plants ion is formed it can attract one additional negatively charged particle; that may be an electron as in the above case, or it may be an anion equivalent (A~). The A~ may denote any mineral anion. As soon as ferrous iron is oxidized by giving off an electron and becomes trivalent, it can bind an A~. That implies that there may be an exchange on the iron atoms of electrons for anions. The iron loses one electron and attracts one anion. To explain the transport of the anions through the cytoplasm, Lundegardh assumes that there are tracks or bridges of cytochromes arranged in transverse direction across the cytoplasm between its outer boundary, called the o-level, and the inner boundary, the /-level, which may be the toiioplast or central parts of the organ. Along these tracks the anions are moved by successive exchanges for electrons. At the o-level the anion is attracted by a ferri-cytochrome. This then takes up an e~~ from an inner cytochrome, loses its anion, which is caught by the second cytochrome, and so on. In this way the anion moves from one iron atom to another through the system. Lundegardh suggests that there are waves of electrons going from the i- to the o-level and, if such a transport of electrons occurs, anions can be moved in the opposite direction. At the /'-level the anions are released from the cytochromes and given off to the vacuole. It is of course impossible, for electrostatic reasons, for a single anion to be released unless it be combined with a partner having the opposite charge, a cation. The anions are thus combined with cations at the /-level. The cations may be hydrogen ions or more likely metal ions, which are abundantly present, and thus the salt becomes stored in the vacuole. The cations can easily be bound to the cytoplasm, because we know, especially from Lundegardh's determinations of the surface potentials of the cytoplasm, that it has a predominantly acid character. It con- tains abundant acid groups with dissociating hydrogen ions, and at these loci, metal cations can be adsorbed by an exchange for the hydrogen ions. I have already mentioned that there are good reasons to assume that in the first step of the absorption the cations really are attached to the plasma by such an adsorption. Lundegardh now as- sumes that the cations are transported through the cytoplasm by re- Hans Burstrom 255 peated such exchanges for hydrogen ions and ultimately reach the /-level. Repeated exchanges of this kind must take place to secure the adsorption equilibria as the cations are withdrawn from the col- loidal system through combination with the released anions at the /'-level. Thus, the driving force for the intake and accumulation of salts is the absorption of anions which is caused by the passage of electrons to the plasma surfaces. The central problem then becomes to ascertain whether we are correct in assuming such a unidirectional wave of electrons. Lundegardh answers this question in the affirmative. The transport of electrons by cytochromes is a well-known part of the iron-catalyzed respiratory system, one end of which is the oxidation by the atmospheric oxygen of ferrous iron carrying one electron in excess. In fact, the oxidation requires that the ferrous iron give off its electron to the oxygen, which combines with hydrogen ions to form water. This is the formation of water which occurs in the respiration. This oxidation, Lundegardh postulates, ought to take place at the o-level, because oxygen is supplied from the external solution, and there must be a falling oxygen gradient from the o- to the /-level. The other end of the respiration, involving dehydrogenations of the respired substrate, is a reduction process and may be restricted to loci of low oxygen tension, which means the /-level. Electrons must be given off there and carried through the cytoplasm to the o-level. We get as a consequence the postulated transfer of elec- trons, and thus the whole mechanism is made possible. The production of electrons is caused by the splitting of the substrate as a normal part of respiration which together with the oxidation at the o-level appears as the salt respiration. In this way it is clear that the absorption of anions is directly connected with the process of respiration. Such a picture involves a fixed polar organization of the cytoplasm in accordance with the oxygen gradient, for which we have no direct evidence. It is absolutely necessary, however, for any theory of ion absorption to assume a polarity within the cell, and this one seems to be both very simple and well founded. The ultimate controlling factor of the absorption and accumulation of ions is then the oxygen gradient in the cell. The theory of Robertson is in its main features very similar to that 256 Mineral Nutrition of Plants of Lundegardh, on which it is largely based. Robertson assumes the same function of the cytochromes, the same polarity, and the same mechanism of combination of anions and cations at the /-level. He has made one change, however, which may be rather important. He does not postulate a successive exchange of anions along a cytochrome bridge and against an electron wave, but assumes that the cytochrome system circulates with the cytoplasm between o- and /-levels. At the o-level, where ferric iron is formed by oxidation, it catches an anion and the whole complex is moved to the /'-level. Here the iron is again reduced by absorbing an electron and the anion is set free. The iron, in the reduced state, returns to the o-level, where it is oxidized, thus completing the cycle. It is worth mentioning that Arisz (/) has pointed out a third possibility, namely, that the transfer of ions between the plasma constituents is conditioned by the continuous breaking down synthesis and of organic compounds, and this means that a high degree of instability of the cytoplasm should be one of the causes of the change in position of the mineral ions. At present we cannot decide which of these two or three assumptions is correct, but this is certainly an important question, because it may contribute to an explanation of the quantitative relationship between the salt respiration and the salt accumulation, or, which is the same thing, the efficiency of the salt respiration. It is quite certain that there is no stoichiometric relationship between salt respiration and the amounts of ions absorbed. Robertson has computed, however, the maxi- mum amount of ions that can be absorbed per unit of respiration if the theory is correct. This is very simple since one electron corresponds to one equivalent of anions and one molecule of oxygen corresponds to four electrons. Thus, one molecule of oxygen consumed in respira- tion may cause the absorption of, at most, four equivalents of anions. His experimental data show that the highest ratios of ion equivalents accumulated to oxygen consumed for storage tissues and excised roots are between three and four. This is no proof of the correctness of the theory, but is quite consistent with it. In a very recent paper Lundegardh (5) has shown, however, that in intact roots this quotient does not exceed 1, which is explained by a respiration connected with an internal transport of salt in the normally functioning roots. In practice, it is Hans Bur strom 257 not the same thing to work with a histologically highly organized root as with a uniform piece of storage parenchyma. So much for the theories. They are obviously speculations to some degree, but, nevertheless, in conformity with all or most experimental data and also with our knowledge of respiratory mechanisms. Several points need to be further clarified, however. The first question is, as already discussed by Robertson, why this special respiration system does not work in the absence of external salts. All integral parts are present in the cell: substrate, enzyme, and oxygen. Here we can only guess. We may refer to an activation of the system by the anions, but that does not explain anything. A second point is, how is it possible to get unequal absorption of anions and cations? If the anions are absorbed in excess, the problem is simple. The excess is combined with hydrogen ions and the vacuolar sap is acidified. The cell possesses no means of preventing such a decrease in pH. Usually there is, however, an excess accumulation of cations, because the rapidly absorbed anions, notably nitrate, are assimilated to a large extent in the cytoplasm and an excess of cations appears in the vacuole. These two questions are probably intimately connected. The second one has been partly elucidated by the work of Ulrich (7) and the author (2). Ulrich showed that the excess of cations in his material was balanced by a production of malic acid which was equivalent to the excess cations, with some exceptions. Our observations showed that this equilibration without exception takes place not at the point of in- take of the ions but at the point of their accumulation or, with the adopted terminology, at the /-level. When an excess of cations is re- leased to the vacuole, it is accompanied by an equivalent amount of malic acid or, at the prevailing pH, of malate ions. As to the source of this malic acid, it is almost certainly formed as an intermediate in the acid cycle of the respiration. This acid cycle forms the fundamental part of the dehydrogenase systems and is, ac- cording to the picture outlined above, localized to the /-level. In this cycle malic acid may be formed from succinic acid and itself trans- formed into oxaloacetic acid. Of course other transformations are also possible. Here malic acid, which is a weak acid, reacts as the undis- 258 Mineral Nutrition of Plants sociated acid, but it must always be dissociated to a fixed extent ac- cording to the prevailing pH, malic acid ^ malate . If an excess of cations is added to such a system, this dissociation shifts to the right, and an amount of malate equivalent to the cation excess is withdrawn from the respiration cycle and is apparently excreted to- gether with the cations into the vacuole. If the excess diminishes, the dissociation reaction is reversed and the acid disappears in the respira- tion cycle again. This is a simple matter of equilibria, and Ulrich has shown directly that the formation and disappearance of the malic acid in connection with the ion accumulation is effectuated by the respira- tion system. We may go one step further. If the ferric ions of the cytochromes are able to bind anions, why should they restrict themselves to the mineral anions, why not attract organic anions as well? Robertson has, in fact, raised this question, and Lundegardh (4) has answered it by assuming that in intact roots, a major part of the salt respiration de- pends upon the action of a respiration system combined with organic anions, called "native" anions. If we assume that the ferric atoms are normally equilibrated by malate, or other organic acids formed in the cell, it is easy to explain why an excess of cation accumulation causes malate formation. As soon as a mineral anion enters the cytoplasm, it is attracted by a ferri-cytochrome and there replaces a malate ion, which disappears in the respiration. Then the anion is moved inward and, if it should happen to be assimilated as regularly occurs with nitrate ions, its place is immediately occupied by a malate ion again. At the /-level the anions (mineral ions or malate ions) are set free and combined with cations in the vacuole (Figure 1). It is, of course, not necessary to assume that this role must be played by malate or malic acid, although it is so in the cases investigated so far. In other plants other acids may fulfill the same function, depending on, for instance, the existing pH of the cytoplasm, the dissociation constants of the acids, etc. A very strong argument in favor of the assumption that the dehy- drogenation system, producing both hydrogen and organic acids, is located at the /-level where the accumulation of the ions is regulated, is given by Lundegardh (4) in a very recent paper. He has shown that the bleeding from a root is dependent upon the release of ions to the Hans Barstrbm 259 vessels, which is a reaction at the /-level, an anaerobic reaction. This is inhibited by such agencies as iodoacetate and fluoride, known to attack the dehydrogenase systems or the anaerobic process of glycolysis. If it is true that the cytochrome system may be occupied by organic acids, it follows as a consequence that the salt respiration system might work with the organic acids present in the cell and not only when mineral salts are added externally. That is just what Lundegardh has assumed to occur, and it is supported by other indications in the same O-LEVEL MEDIUM 02 © CYTOPLASM I- LEVEL VACUOLE ^RESPIRATION *■ oxiDAT/OH SYS T E M © A CYTOCHROMES DEHYDRO- 1 Cif/V/*770/V H acid cycu\ Fe Fe ORGANIC ACIDS Fe y ■« (e t PFI (A) (org.-A) »(A)>»---(ORG.-A)^HZE(ORG.-A^ v(g^(ORG-A-)>» Figure i. Scheme for the active absorption of ions. The salt respira- tion system is oriented transversely across the cytoplasm. At the inner, dehydrogenation level, protons and electrons are produced, and they are moved to the outer, oxidation level in exchange for salt cations and anions, respectively. Solid arrows denote only chemical reactions or equilibria; broken arrows, mechanical transport. direction. For instance, Robertson has shown that a respiration sensi- tive to hydrocyanic acid also occurs in the absence of mineral salts under certain conditions. Inhibition by hydrocyanic acid may cut off not only the whole additional salt respiration but a part of the apparent ground respiration as well. These and other observations show that the cytochrome-regulated respiration is not exclusively dependent on the addition of salt. It is apparently only a question of relative rates of res- piration with and without mineral salts. Salts, presumably anions, greatly activate a respiratory system, and it may be assumed that this activation involves the exchange of the organic acids for the mineral 260 Mineral Nutrition of Plants anions. This means that the cytochrome system saturated with organic anions should be much less active than when it contains attached in- organic anions. The function of the cytochrome system depends upon its ability to carry electrons between the iron atoms. If the latter bind anions, the activity must depend upon the rate of exchange of the elec- trons for the anions, a reaction which certainly must vary with the kind of ion. The fact that large bivalent organic anions are much less ex- changeable than the inorganic ions, might explain the activation of the respiration system by the added mineral salts. In this way the organic acids occupy a central position of interest. Their role is, however, far from clarified and can hardly be tackled from common physiological points of view. If this theory gives a true picture of the mechanism of the ion absorption, biochemical studies are necessary for further progress within this field. There are still other difficult questions to answer. One is the real cause of the specificity of the ion absorption: why do not all plants absorb all ions in the same proportions, for instance always potassium more rapidly than sodium, as was to be expected from the physical properties of the elements and their importance for the exchange phe- nomena regulating the absorption. It is obvious that the recorded theo- ries can only claim to explain the principles and the fundamental mechanism underlying the absorption of ions, but there are details which hitherto have not been seriously considered at all. I do not think, however, that such gaps in the theories detract from their value as a solid working basis for explaining the active absorption of ions. REFERENCES i. Arisz, W. H., Proc. Koninkl. Nederhwd. Akad. Wetenschap., 50: 3 (1948). 2. Burstrom, H., Ar\iv. f. Bot., 32 A, Nr. 7 (1945). 3. Lundegardh, H., Arkjv . f. Bot., 32 A, Nr. 12 (1945). 4. •, Ann. Roy. Agr. Coll. Sweden, 16: 339 (1949). 5. •, Ann. Roy. Agr. Coll. Sweden, 16: 372 (1949). 6. Robertson, R. N., and Wilkins, M. J., Austr. J. Sclent. Res., B r: 17 (1948). 7. Ulrich, A., Am. J. Bot., 29: 220 (1942). 8. Wanner, H., Ber. schweiz. botan. Ges., 58: 383 (1948). CHAPTER \\J The Translocation of Minerals in Plants* O. BIDDULPH I n multicellular organisms where one organ specializes in mineral absorption, a distribution system for those minerals is an inevitable development. The efficiency of the distribution system will, in general, reflect the over-all efficiency of the organism since the growth rate is dependent on an equalization of the necessities for growth between the various organs and tissues. The growth rate may not outstrip the supply of material upon which it is dependent. In plants there is no concentrated expenditure of energy for either maintenance of body temperature or for muscular contractions, hence, in keeping with the general level of specialization, there is no need for a highly developed "circulatory system." Nevertheless, the transloca- tion system must be efficient enough to distribute rapidly the incom- ing materials, i.e., minerals, and to equalize concentrations between various parts as tissue elaborations proceed. Even under the conditions of reduced growth rates there is a continual exchange of materials within the physical structure of the protoplasm which demands a con- stant supply and removal of the basic components. The cellular entities, such as the major morphological protoplasmic constituents, maintain a relatively constant pattern even though the molecular or atomic con- stituents undergo continual replacement. It is my purpose here to review the translocation of minerals in plants with a view to correlat- ing some of the observations on translocation with the requirements of the organism. It will be necessary in many instances to resort to #The writer gratefully acknowledges the active participation of J. Witt, J. Rediske, and C. Woodbridge in the research upon which parts of this report are based. Figures i, 5 and 8, and 4 and 7 are from their respective theses in the Department of Botany at the State College of Washington. 262 Mineral Nutrition of Platits generalizations. I realize that generalizations can be dangerous, yet it is through them that progress is made. For the purpose at hand, I shall attempt to travel that narrow path between oversimplification on the one hand and undersimplification by too ardent a treatment of minutiae on the other. TRANSLOCATION ACCOMPANYING GERMINATION The translocation of minerals from the storage regions of the seed to the developing plumule, hypocotyl, and radicle constitutes the initial mineral translocation in the ontogeny of the developing plant. Accom- panying the hydration of the seed there is a solution of the minerals present. Most seeds contain sufficient minerals to serve the developing seedling until the roots, stems, and leaves are well developed. During the germination of red kidney bean seeds in moist Scot towels, we have observed a net loss of iron from the seeds amounting to approximately 50 per cent of the total initially present. This amount of iron was gained by the toweling at the expense of the seeds and, therefore, was lost to the seedlings as they were removed to other media after two days. At this stage the roots were not sufficiently developed to reabsorb the iron from the toweling. The major portion of the iron remaining in the cotyledons was readily translocated to the developing plumule, hypocotyl, and radicle. This translocation parallels closely the translocation of carbohydrate (Figure 1). These results are in good agreement with the movement of nitrogen, phosphorus, magnesium, and potassium in germinating wheat seeds as reported by LeClerc and Breazeale (5). This is true even though the storage materials in wheat are found chiefly in the endosperm while in beans the reserves occur in the cotyledons. Studies were made in both normal daylight and total darkness. Translocation of iron, as well as carbohydrate, proceeds much more slowly in darkness than under conditions of normal daylight (alternat- ing night and day). Both Whitmore (10) and Withrow (//) have re- ported somewhat similar results. It is clearly indicated that light exerts an influence on the rate of translocation of both reserve foods and iron from the cotyledons into the developing plant axis. The mechanics of this efTect is unknown at present. O. Biddulph 263 x x Dry matter * Iron *" '"< £ 3 4 5 6 Days of Germination Figure i. The dry matter and iron content of bean seedlings during the first seven days of germination expressed as percentage of material in un- germinated seeds (less seedcoat). Graphs for plants kept in total darkness are included for the fourth to the seventh day only. TRANSLOCATION OF MINERALS IN THE XYLEM Translocation in the mature plant is a complex phenomenon. It is dependent on the initial absorption processes carried on by the epi- dermal cells of the root. From this point there begins a cell-to-cell 264 Mineral Nutrition of Plants transfer across the cortex to the cells surrounding the xylem. Here, in a process "akin to secretion," according to Hoagland (4), the minerals enter the xylem system wherein they ascend to the aerial extremities of the plant. This process of secretion, if it may be so called, is one of the least known of all processes in the realm of mineral translocation. Its elucidation undoubtedly must await a fuller understanding of the absorption process. Since so little is known of this "secretion" process as it applies to translocation, it will not be discussed here. The delivery of mineral elements to the leaves of mature plants most certainly occurs primarily through the xylem tissues. Stout and Hoag- land (9) have contributed the most direct evidence bearing on this point. In their experiments the phloem was isolated from the xylem by the insertion of waxed paper between the tissues but the paper left the longitudinal continuity of both elements intact. Radioactive nutrient elements were then placed in the root environment and the path of these elements was followed by the radiation emitted as the nutrients ascended the stem. At the place where xylem and phloem had been separated a preference of paths was shown: namely, most, if not all, movement in an upward direction occurred in the xylem. This experi- ment seems direct and conclusive for conditions where at least moderate transpiration exists. The fate of minerals as they are swept upward by a rapidly moving transpiration stream might be varied according to the mineral element, the plant, and the conditions existing within the plant at the moment. There are several possibilities, enumerated as follows: 1) a portion of the material will be captured by the cells adjacent to the xylem, in particular the cambium and young phloem; 2) a portion may move laterally via rays and the like to actively metabolizing cells; 3) a por- tion may be deposited in the leaves having moved there via the transpi- ration stream; or 4) a portion may move directly to the apical primordia and adjacent regions of active metabolism. In a general sense there are two basic phenomena which influence the direction of movement of minerals within a plant. These are metabolic use and transpiration. The intensity of these factors in any tissue will determine the net movement to the tissue. The metabolic use of an element establishes gradients responsible for a continued flow. O. Biddulph 265 Transpiration delivers to xylem extremities dissolved materials in proportion to the amount of water lost from the tissue. It should be remembered that the capture and metabolic use of elements from the transpiration stream may continually alter the mineral composition within the transpiration stream as it moves upward and, hence, the mineral composition of solutions delivered to the extremities of the xylem system may differ from that which begins its movement in the root region. A characteristic distribution of radiophosphorus (P32) to leaves of all ages on bean plants is shown in Figure 2. It can readily be seen that uniformity of delivery is more characteristic of older leaves than of younger ones. We have been unable to assign relative weights to the two factors, transpiration and metabolic use, in the determination of the quantity of P32 which will move to a given tissue, but it can be seen that the highest relative concentration invariably occurs in the region of highest growth rate. A carefully controlled microclimate around various meristematic tissues might aid in the further solution of this problem. DEPOSITION OF MINERALS IN LEAVES So far we have assumed that no chemical or physical reactions occur in the xylem system which might precipitate or absorb mineral com- ponents of the transpiration stream. I wish to discuss this possibility and to present positive evidence bearing on this point. Precipitation reactions in the xylem would alter the composition of the mineral solu- tion available at the xylem extremities. Nutritional unbalance may then result. This condition could be further accentuated by a more rapid re- moval of water than of minerals (all or certain ones) from the trans- piration stream. Olsen (8) has presented some evidence that there may be a precipita- tion of ferric phosphate in the veins of leaves of the corn plant under certain growth conditions. This seems entirely possible. The phenome- non occurs on root surfaces where the epidermal cells begin the initial accumulation of minerals from the nutrient medium. Here, actively metabolizing cells are acquiring materials from a solution surrounding them. These materials cross the cortex and are "secreted" into the xylem 266 Mineral Nutrition of Plants 2000- 1750- 1500- o E 4> C 3 o o 1250- 1000- 7 50 CM °- 500- 250- 12 3 4 5 6 Position of Leoves on stem ( base : tip = left irighP Figure i. The amount of radiophosphorus acquired plotted against the position of the leaves on the stem for ten bean plants grown in the same nutrient tank. Oldest (opposite) leaves on extreme left; youngest leaves on right. O. Biddulph 267 where they ascend via the transpiration stream to the xylem extremities. At these extremities the actively metabolizing cells acquire minerals from a solution which they in turn surround. In other words, two separate absorption processes must occur before the mesophyll cells of leaves acquire minerals, one by the cells at the root surfaces and the other by the cells surrounding the xylem. There is no valid reason to assume that precipitation reactions would not occur in the xylem ex- tremities except perhaps that they may be partially protected by precipi- tation reactions occurring at the root surfaces and so may be partly "screened" from the possibility of reactions of similar intensity. Even so, precipitation reactions can be duplicated in both places. Our evidence is presented as follows. Exactly comparable studies of precipitation reactions at root-solution and vein-mesophyll junctions cannot be made for anatomical reasons; therefore, two methods of study are necessary. The root-solution junc- tion has been studied by conventional tracer methods, while the vein- mesophyll junction has been studied by autoradiography. Results are limited to iron and phosphorus. The uptake and translocation of iron by bean plants was measured by the usual colorimetric methods. The roots were not washed prior to analysis and so contained most of the iron which had accumulated both in the roots and on the root surfaces during the growth period. Figure 3 (lower part) shows the iron con- centration of bean roots when grown for 12 days in solutions of various iron concentrations. The upper part of Figure 3 shows the distribution in the leaves of radioiron which was placed in the nutrient solution during two additional days of growth. These data show that the pres- ence of a large amount of precipitated iron on the root surfaces has inhibited the uptake of additional iron (Fe55-59). The composition of the precipitate on the root surfaces has been analyzed and is known to be predominantly ferric phosphate at pH's below 6.0. At pH 6.0 some calcium is present while at pH 7.0 calcium is an important constituent along with the iron and phosphorus. In the nutrient solution used, the phosphorus-to-iron ratio was approximately 10:1. With regard to the entry of phosphorus into the plant, the following experiment will show that it is influenced by the presence of iron in the nutrient solution. In this study the phosphorus-to-iron ratio was ap- 268 Mineral Nutrition of Plants proximately 1:1. Figure 4 shows that as iron and, simultaneously, phos- phorus increase on the roots, movement of phosphorus into the top is impaired. This, with the foregoing experiment, furnishes substantial evidence that precipitation reactions may occur on roots which may in- fluence the ready entrance of iron and phosphorus into the root, and, therefore, the plant as a whole. 4 th. leof 3rd. leaf 2 nd. leaf st. leaves :e/g.EM. Roots ] 1 1 ] 0 25 50 0 25 50 1 25 50 1 25 Mq. 1 0 1.0 0 1.0 0 1.0 0 0.5 mg.Fe/g. D.M. Iron concentration of nutrient media below: I.Op.p.m.Fe O.lp.p.m.Fe O.OIp.p.m.Fe 0.002 p. p.m. Fe Figure 3. The amount of radioiron moving into the various leaves on four bean plants each grown at different iron concentrations and having dif- ferent amounts of iron associated with the roots. The most direct evidence that mineral nutrients may concentrate in veins and become immobilized has been attained for iron. Fe5° auto- radiographs on no-screen X-ray film furnish a particularly good medium with which to show this phenomenon. Autoradiography were prepared of bean plants grown at (a) pH 4.0 and medium phosphorus (0.0001M), (b) pH 7.0 and medium phosphorus (0.0001M), and (c) pH 7.0 and high phosphorus (0.001M). Condition a results in healthy green plants. Condition b yields plants with chlorotic mesophyll, later becoming generally chlorotic. (The autoradiographs were made at the time that the tissue adjacent to the veins developed chlorosis). Condition c results in general chlorosis throughout the leaf (Figure 5). The distribution of Fer'5 in leaves under the above conditions shows: rapid entry and equal distribution throughout at pH 4.0 and medium to > o c o o o o a. o .c Q_ C a o> o> o> o c a> o a> Q. 20- O. Biddulph % gain in P. Fe cone of tissues 0 te£ves Leaves 269 (A 8 * o 0) C O a> 4 .2 c " E o t c c ' 2 sq. in. 0.886* Shoot length X average area per leaf of leaves > 2 sq. in. °-525 * "■Significant at i per cent point or beyond. 282 Mineral Nutrition of Plants 1942 1943 194? Figure i. Average annual yield per tree in a New York Mcin- tosh apple orchard under three levels of nitrogen fertilization. Annual treatments: H trees, 1.5 lb. N; M trees, 1.0 lb. N; L trees, 0.5 lb. N. to the weather conditions in the period of bloom. In the six years, 1944 was the only one in which there was a long period of ideal weather for pollination, and no frost. In 1945, crop failure was the result of frost and cool rainy weather during the entire period of bloom. The years of 1946 and 1947 were less favorable for pollination than 1942 and 1943, although there was some opportunity in all four years. Thus, weather during the bloom was the most significant factor limiting yield in this orchard. In 1943, 1944, and 1946 the total yields of the low nitrogen (L) trees were significantly less than those of the intermediate nitrogen (M) trees. The direct causes of the decreased total yield due to nitrogen de- Damon Boynton 283 ficiency were decreased bloom and set, and smaller size; all of these were probably the result of the decreased leaf surface due to rather ex- treme nitrogen deficiency. Under the circumstances, even though no increase in the amount of fruit grading as fancy occurred on the aver- age due to the first increment of nitrogen, this increment was justified. Had conditions been favorable for set in 1945, tne ngnt bloom which occurred on the low nitrogen trees in that year would undoubtedly have caused a markedly smaller crop than that on the M and H (high nitrogen content) trees. There is no such justification for the second in- crement in nitrogen. While there was growth response to it, the added growth was never associated with total yield response and it was always associated with decrease in fruit color. Figure 2 shows that there was also fluctuation from year to year in the color of the fruit samples. It should be noted that there was a nitrogen effect on fruit color even in 1942 when there was no effect of nitrogen on yield, and that in 1944, the year of greatest crop, the average fruit color was poorest. The fluctuations from year to year „ o 1942 1943 1944 1946 1947 Figure 2. Average percentage of fruit samples having fancy color in a New York Mcintosh apple orchard under three levels of nitrogen fertilization. Annual treatments: H trees, 1.5 lb. N; M trees, 1.0 lb. N; L trees, 0.5 lb. N. 284 Mineral Nutrition of Plants seem to have been related to the weather conditions at the harvest period in the years of moderate crop. Cool bright weather at harvest time was conducive to the best color development. In addition, the very heavy crop in 1944 may have operated to reduce fruit color develop- ment in that year. PRACTICES AFFECTING NITROGEN SUPPLY These data should serve to illustrate some of the complexities of the problem of controlling nitrogen effects on apple trees with the purpose of obtaining maximum production and fruit quality. The growth phenomena that are most important in determining yield occur largely in the early part of the growing season while fruit color and quality develop at the end of the growing season. Thus, a relatively high nitrogen status of the tree, favoring high yield, is de- sired in the early growing season, and a relatively low nitrogen status, favoring fruit color and quality development, is desired in the latter part of the growing season. In the northeastern United States, weather most often is a dominant factor that limits crop size and quality. The grower cannot control the weather; but he can manipulate certain cultural factors, the most pertinent of which are 1) nitrogen application — material, rate, time, and method; 2) ground cover; and 3) pruning and spraying practices.* Nitrogen application Readily available inorganic or simple organic forms of nitrogen are used for ground application in Mcintosh apple orchards in prefer- ence to forms which release available nitrogen over a long period of time. When moderate applications of such materials as ammonium nitrate, sodium nitrate, or ammonium sulfate are made in the fall or early spring, most of the soluble nitrogen is usually absorbed by plant roots before midsummer. At high rates of application, and with sparse ground cover or under heavy mulch, considerable available nitrogen may remain in the soil throughout the summer. Under such circum- *In the ensuing discussion it is assumed that nitrogen is the only limiting nu- trient, that soil moisture and aeration are satisfactory, and that no unusual condi- tion such as winter injury, trunk girdling, or insect or disease troubles have taken on major proportions. Damon Boynton 285 stances it is not possible to control the nitrogen effects. Even when rather low rates of nitrogen application are made, the period of time that soluble nitrogen persists in the soil will vary greatly with weather conditions. Heavy winter rainfall following fall fertilization may in one year cause all applied nitrates to leach beyond the rooting zone, whereas following a winter of low rainfall all the nitiates applied dur- ing fall may be available for absorption. Because of this fact and because occasionally serious winter injury to the trunks of trees occurs follow- ing fall fertilization, it has been customary to apply nitrogen fertilizer in the early spring in New York. Late spring and summer applications are not made because they may inhibit fruit color formation and do not furnish available nitrogen at the time it is used in vegetation, fruit set, and initiation of flowers for a subsequent crop. The most common method of ground application has been to broad- cast in a ring under the branches of the trees. This has developed as a result of the fact that there is less use of the fertilizer by grass if it is applied in a concentrated band than if it is broadcast over the entire orchard floor. There has been little critical evaluation of this method of application, however, and it is possible that it is not the best under all circumstances. Certainly, the grass cover over the orchard floor deserves some attention and it may be worth while to fertilize it occasionally if not annually. The use of urea sprays has recently attracted attention as a method of fertilizing Mcintosh apple trees. Originally suggested in 1942, it has been employed commercially by apple growers rather widely during the past season. Table IV summarizes the results of some comparisons between ground fertilization and urea sprays made on Mcintosh apple trees in 1948. These and other studies indicate that urea spraying is at least as effective as equivalent rates of ground application, in terms of yield, growth, and fruit color and size. The ultimate value of this method of nitrogen fertilization will probably be determined by the efficiency of direct absorption by apple leaves. Preliminary studies indi- cate that half the urea adhering to the leaf surface may be absorbed by the leaf within a few hours of application. Since as much as half of the spray material applied to the leaf may drop off, it is possible that frequently no more than a quarter of the urea applied to the leaves 286 Mineral Nutrition of Plants TABLE IV Comparison of the Effects of Sprays and Spring Ground Applications of Urea on the Yield, Growth, and Fruit Characteristics of Mcintosh Apples in 1948. Average of Three Experimental Plots in Western New York Av. Per- centage Wt. Av. Spurs Urea Treatment Av. of Fruit Fruit Termi- Setting Total Samp- Sample nal Fruit | N Methods of Yield ling (60 Growth f Pounds Application per Grading fruits) per tree plot Fancy 0 lb. 0.6 1.2 2.4 1.0 !-3 Ground, April Ground, April Ground, April Ground, April 13.9 bu. 14.6 16.2 18.5 82.8 66.3 56-3 40.4 14.6 lb. 15.4 16.2 16.6 18.5 cm 23-i 26.4 28.5 26.2% 27.1 311 321 Spray Calyx, 1st, 2nd 19.0 59.0 16.3 26. 1 32.6 covers 2 pink, calyx, 1st cover 17.7 SS-6 16.6 28.0 K.i LD*5 per cent 1.8 8.5 0.6 2.6 2.0 *Least difference for statistical significance at odds of 19 to 1. "("Twenty shoots per tree were measured at random after growth had stopped. jThe percentage of bloom and set on at least 200 vigorous spurs on 4 sides of each tree were determined. Since practically all spurs bloomed in 1948, the percentages are based on the total number of spurs counted. enters them, the rest going to the ground where it is ultimately absorbed by roots. There appears to be far less efficiency in absorption by the upper surface of an apple leaf than by its lower surface, and young leaves seem to be more efficient than older ones. So it remains to be seen whether or not this development can be used to improve our con- trol of the nitrogen nutrition of apple trees. Ground cover The presence of ground cover may be indirectly important in per- mitting control of nitrogen effects on apple trees. Nonleguminous sod uses large quantities of nitrogen in its growth and can be relied on to Damon Boynton 287 remove soluble nitrogen from the soil solution during the latter part of the growing season. This means that fertilizer practices in apple orchards should be designed to encourage grass growth. In New York apple orchards, grass responds most often to nitrogen and liming. Usually sparse grass growth in the areas between the trees is due to lack of nitrogen. Failure of cover growth from lack of calcium and magnesium occurs under the trees where acidification by sulfur sprays has been most rapid. While there is no experimental substantiation of this idea, it is probably worth while from the standpoint of nitrogen control alone to fertilize for healthy grass growth in Mcintosh apple orchards. The use of supplementary mulching materials spread under apple trees is a common practice in many New York apple orchards. When these mulches are composed of materials high in carbon, like straw or nonleguminous hay, their initial effect is to depress the available nitrogen supply in the soil. This effect may be apparent for a year or more in apple trees mulched with high carbon materials. If the mulch blanket is maintained over a period of years, ultimately nitrates will build up under it and will remain high throughout the entire year. Thus, heavy mulching even with nonleguminous materials causes a loss of control of nitrogen effects. There has been no critical study of moderate mulching in relation to nitrogen fertilization, but it is prob- able that if annual mulch additions are light enough so that grass grows through the blanket freely, satisfactory nitrogen control may be maintained. Pruning and spraying practice Pruning and spraying practices may have important indirect in- fluences on the nitrogen requirements and responses of apple trees. Heavy pruning has been repeatedly demonstrated to act in the same way as nitrogen fertilization, and in orchards where pruning is severe, the nitrogen fertilizer program may be somewhat curtailed in order to improve fruit color without loss of productivity. Nitrogen fertilization, to a considerable extent, has been used as a means of overcoming the detrimental effects of fungicides on vegeta- tion of apple trees. It was a common experience for apple growers 288 Mineral Nutrition of Plants changing from lime-sulfur to elemental sulfur fungicidal spray pro- grams to find that it was desirable to reduce the rate of nitrogen ferti- lization. Recently, an experiment reported by Dr. D. H. Palmiter has indicated that elemental sulfur sprays may also reduce vegetation and that nitrogen fertilization may offset these effects. He found that over a six-year period apple trees sprayed with Fermate (ferric dimethyl dithiocarbamate) and given no nitrogen fertilizer yielded as much as trees sprayed with elemental sulfur and given 5 pounds of urea annu- ally. Both of these treatments resulted in an increase in yield of 38 per cent or more over the yield of trees unfertilized and sprayed with sulfur. While there was some nitrogen in the Fermate spray, analysis of leaves indicated that the increase in yield could not be attributed to that; rather, it seems to be due to absence of spray injury. OTHER CONDITIONS AFFECTING NITROGEN RESPONSES There are, of course, a large number of conditions that may limit nitrogen responses besides the ones that have been discussed. In New York Mcintosh orchards, excess or deficiency of soil moisture and serious injuries to the leaf surface from insects, disease, or spray ma- terials are not uncommon limiting conditions; occasionally potassium and magnesium deficiencies may be significant factors. Their effects on nitrogen responses are of several sorts. Excess moisture in heavy soils results in poor soil aeration which may both reduce the absorbing sur- face of the root system and the ability of the absorbing surface that is there to function efficiently. This will ultimately lead to the death of the tree. Most commonly, the periods of poor aeration occur in the spring and are short enough to permit survival of mature trees in marginal locations; the effect is one of general devitalization and failure to respond to good management in growth and productivity. Soil moisture deficiency, occurring usually in midsummer or later, may have no influence on growth or set of fruit, and may actually result in increased bloom the following spring. However, the size of the fruit in the year of drought is decreased in proportion to the duration of the period of moisture unavailability in the soil. Fruit color may be some- what improved as a result of some moisture deficiency but, if the deficiency period is prolonged and continues until harvest, the quality Damon Boynton 289 of the color may be poor and preharvest drop tends to be very heavy. Injuries to the leaf surface from insects, diseases, spray materials, or mineral deficiencies modify nitrogen responses in different ways, de- pending on the time and severity of loss of effective leaf area. Early damage or defoliation such as that caused by severe primary infections of apple scab or lime-sulfur injury, may reduce vegetative growth, fruit set, and flower initiation for subsequent crops. Later leaf injuries, such as those caused by mites, by leaf hoppers, by arsenical injury, or by potassium or magnesium deficiencies may have little or no influence on growth, set, or subsequent crop, but tend to decrease fruit size and quality and to predispose the trees to heavy early fruit drop. DIAGNOSTIC OBSERVATIONS ON NITROGEN RESPONSE These, then, are some of the main practices which the grower may manipulate in the attempt to control nitrogen effects on the Mcintosh apple tree and some of the other conditions that influence nitrogen responses. Under the variable weather conditions of the northeastern United States, it seems unlikely that very close control will be possible but some measure of it may be obtained by adjusting practices accord- ing to careful diagnostic observations. Because of the perennial nature of the apple tree, these diagnostic observations need to take into ac- count separately : tree responses to conditions of the current season and tree responses to conditions of previous seasons. The first and most important observation to be made is an evaluation of the crop on the tree, in relation to the bearing capacity of the tree and the bloom, and in relation to the best average fruit quality that is practi- cally attainable. Next, measurements of shoot growth, leaf size, and leaf color appear to be particularly useful in diagnoses of this sort. These are interrelated more frequently than not but furnish in com- bination useful quantitative evidence of the nitrogen status of an apple tree. Table V illustrates the effects of four rates of nitrogen fertiliza- tion on some of these measurements on shoots from two New York Mcintosh apple orchards. For such measurements to be useful under field conditions, they have to be easily made and to be referred to a set of growth standards. Leaf color may be determined by reference to color standards for apple leaves now available or may be made from 290 Mineral Nutrition of Plants TABLE V Average Leaf Number, Leaf Color, Leaf Size, and Shoot Length of Representa- tive One- Year-Old Terminal Growths from Two Mcintosh Apple Orchards Under Differential Nitrogen Fertilization in 1948 Average Total Average Area N Applied Average Leat Number per Leaf Annually Average Leaf Shoot for Leaves > (sq. cm.) of 4 per tree Chorophyll Length 2 sq. in. Largest Leaves mg./65 sq. cm. 0.0 lb. 1.42 16.5 cm. 10.2 32-7 0.6 1 75 21 .2 11. 7 35-5 1 .2 1 .96 28.4 *4-7 42.5 2.4 2.32 31.2 16.0 41 . 2 alcohol extracts of leaf disks. Leaf size is easily and accurately deter- mined by measurement of length and width and by reference to con- version tables previously prepared. These measurements may be rated on the basis of reference groupings like those in Table VI. Terminal growth of shoots and development of leaves in number and size are determined to a considerable extent by the food reserves in the tree at the beginning of the growing season. Leaf color, on the other hand, reflects more closely the nitrogen supply in the tree at the time of observation. Thus, evidence of the effects of food reserve and of the current nitrogen supply may be obtained from these four TABLE VI Tentative Growth Standards for New York Mcintosh Apple Shoots and Leaves Standard 1 (low) 2 3 4 5 (high) Shoot length, cm. < 15 15-20 20-25 25-30 > 30 Number leaves > 2 sq. in. per shoot < 10 10-12 12-14 14-16 > 16 Size largest leaf, sq. cm. <3« 3°"35 35"4° 40-45 > 45 Leaf color (July 15- Aug. 15) Chlorophyll per 65 sq. cm. surface < 1.6 1. 6-1. 8 1 . 8-2 . 0 2 . 0-2 . 2 > 2.2 Mcintosh leaf color 1 2 3 4 5 and > standards Damon Boynton 291 measurements. If shoot length, leaf numbers, leaf size, and leaf color are all abnormally low or abnormally high, it is likely that both the food reserve and the current nitrogen supply were abnormally low or high. If growth, leaf numbers, and leaf size are low and leaf color is high, it is likely that food reserve was low but that the current nitrogen supply has been increased. If growth, leaf numbers, and leaf size are high and leaf color is low, then it is likely that the food reserve was considerable but the current nitrogen supply is not great enough to maintain color. Determination of leaf color at intervals throughout the growing season may give additional diagnostic evidence of the course of current nitrogen supply. Figure 3 shows the chlorophyll trends and total nitrogen trends in leaves from a Mcintosh apple orchard under six nitrogen treatments, involving different levels, times, and methods of application. In all cases the midseason peaks of the curves are de- termined by the total rate of nitrogen application: the higher the rate, the higher the peak. The peak for treatment C was reached later than for the other treatments; this was associated with the fact that three of the six urea sprays were applied after the middle of June. All of the chlorophyll curves tended to flatten at the end of the growing season more than they do in many years. This may be due to the fact that the orchard was harrowed several times in July in order to eliminate deep ruts created by the heavy spray machinery, a practice which may have released nitrate in midsummer and checked the normal sod cover at the end of summer. Yields were low in this orchard due to poor pollination and there were no significant differences among them, but fruit color was inversely related to nitrogen in all cases. It would seem, from this and other studies, that the seasonal chlorophyll curve as- sociated with highest yield potential and best fruit color would be rather high in early summer and rather low at harvest time. The leaf color range of 1.8 to 2.0 mg. per 65 sq. cm. surface, in July, reflects a compromise between these two opposing objectives. Of course, interpre- tation of such measurements must be based on experimental nitrogen response studies over a period of years in the climatic zone where the diagnosis is needed. This is because the nitrogen responses of an apple variety differ in different zones as well as from year to year. Whereas the range of group 3 in Table VI seems to be correlated with satis- 292 Mineral Nutrition of Plants CHLOROPHYLL 2.6 2.4 2.2 ^ 2.0 a> 2T 1.8 a - 1.6 c cadmium, and mercury. These results, while confirming the indispen- sability of zinc and copper in amounts greater than those found in the nutrient medium, were interpreted as permitting no final conclusion as to the role of cadmium, lead, and mercury. The possibility that these elements, or others studied by a similar technique, may be required in amounts smaller than the incidental impurities which could not be re- moved from culture solution by the present technique cannot be a priori excluded. If these views are accepted there can be no objection to regarding almost every element in the periodic table, and particularly those most frequently encountered in plant tissues, as susceptible of being shown at some time to be essential for plants. What can be asserted definitely is that, if an element now regarded as dispensable for a given plant should at some future time be found essential, it will be shown to be required in exceedingly small amounts — within the limits of con- tamination still encompassed by the refined methods now used for purifying the nutrient medium. This quantitative approach to the problem of essentiality of micronutrients is regarded not as a mere theoretical generalization but as a point of view conducive to a search for more refined analytical methods and procedures for growing plants which would make it possible to investigate the status of a number of new elements in plant nutrition. The discussion thus far has dealt with those advances in the explora- tion of essentiality which were made through experimental modification of the external medium. Failure of the plant to grow was taken as the physiological yardstick by which nutrient requirements were measured. Except as a general surmise, experiments of this kind do not help to determine the function which the essential element plays within the plant. Regardless of how many different functions an element may perform within the plant, it is obvious that if the insufficiency of a nutrient resulted in blocking only one crucial reaction, growth would be arrested. These considerations suggest an alternative approach to the problem of essentiality of inorganic nutrients. Rather than measuring the effect of the removal of an element from the external nutrient medium on 326 Mineral Nutrition of Plants growth, would it not be possible to indentify either an essential cellular constituent or a crucial biochemical reaction in which the inorganic element participates? For some of the essential elements the answer to this question has been obvious as soon as the chemical constitution of cellular substances was established. There was no difficulty in assigning an indispensable role to carbon, hydrogen, and oxygen solely on the basis of their entering into the composition of all living matter. As for the elements derived from the soil, the indispensability of nitrogen, sulfur, and phosphorus was adduced at an early period from their identification with proteins and nucleoproteins. As far as cations are concerned, the discovery that magnesium is an integral part of the tetrapyrrolic chlorophyll molecule assured an essential status to that element irrespective of what other functions it may perform in the plant. Calcium combines with pectic acid to form calcium pectate in the middle lamella of the cell wall. This leaves only two elements on the classical list, potassium and iron, to which no essential status was assigned solely on the basis of then- entering into the chemical composition of cellular constituents. With respect to potassium this state of affairs persists to this day. No organic compounds containing this essential element have been detected among the components of plant cells although suggestions have been made that it may combine with proteins. The case of iron deserves special treatment. Although it is known today to be an essential component of cellular constituents, that discovery followed an initial path rather distinct from analytical biochemistry. It resulted from a series of in- vestigations whose primary objective was the understanding of a phy- siological process. In the case of iron the process investigated was respiration. Before embarking on this phase of our discussion, it might be well to state certain premises. A convincing demonstration that a given element is indispensable to some vital process would suffice to establish its essentiality, even in the absence of appropriate growth experiments or corroborative analytical evidence on plant constituents. It is not in- conceivable that for some micronutrient required in exceedingly minute quantity, the previously discussed experimental difficulties would make it impossible to remove completely the element in question either from Daniel I. Anion 327 the nutrient medium or the seed. On the analytical side, the element may enter into the composition of some organic intermediate of such great structural instability or low concentration in the tissue, or both, as to escape detection. It is interesting to note that an approach to essentiality through the study of function was initiated over a half a century ago, in a manner which proved of historic importance to general biochemistry. In 1897 Bertrand (75) in France, reported that manganese was consistently associated with the activity of an oxidizing enzyme in plants, laccase. He came to regard manganese as an essential constituent of the oxidase system and hence essential to plant life. This announcement linked for the first time a metal with an oxidizing enzyme in living cells. It is true that recent work has shown that the effective metal in laccase is copper rather than manganese (31,51), but the principle first proposed by Bertrand has retained its force. The most fruitful development of the functional approach to essenti- ality of micronutrient had to await the evolution of modern biochemi- cal techniques. The small amounts in which these nutrients were re- quired by plants pointed very early to their probable catalytic function and, as first suggested by Bertrand (75), their association with enzymes. But the experimental proof for this hypothesis came much later. For boron, manganese, and molybdenum, we lack to this day precise formu- lation of their biochemical function in the plant. Their indispensa- bility was established by growth experiments. Even for the others to which recent research has assigned some biochemical function, their essential status was already known from growth experiments. Thus the history of the micronutrients reveals that important physiological advances and the agricultural application of our knowledge of the indispensability of iron, boron, manganese, copper, zinc, and molybde- num occurred either in advance, or in the absence, of any knowledge of their functions within the plant. The biochemical approach, however, has led to important advances in recent years. The study of respiration led Keilin (28) and Keilin and Hartree (29) to the identification of four iron porphyrin com- pounds which constitute the cytochrome system and are essential in the respiration of aerobic organisms. Here then was biochemical evi- 328 Mineral Nutrition of Plants dence of such validity that it would have established the essential status of iron as a nutrient, even if results of growth experiments were not available. Two other enzymes found in plants, peroxidase and catalase, have as their prosthetic groups iron-porphyrin compounds. Another enzyme widely distributed in plants and capable of participating in the respiratory process is polyphenoloxidase. This enzyme depends for its activity on the reversible oxidation and reduction of copper. Copper is the prosthetic group of the enzyme and cannot be replaced by any other metal (•?■?, jo). Other copper enzymes in plants are laccase, noted previously, and ascorbic acid oxidase. In recent years the enzyme car- bonic anhydrase was isolated from red blood corpuscles and was shown by Keilin and Mann (^2) to have zinc as a prosthetic group. Present indications are that carbonic anhydrase also occurs in higher plants (/6). In the examples just cited the metal micronutrient is the prosthetic group of the enzyme. Its place cannot be taken by any other element. In this it fully meets the test of specificity which has been previously designated as a criterion of essentiality in growth experiments. It was already implied, however, that the identification of an element with a specific function in no way excludes other roles which the element may perform. An excellent illustration of this principle is the well-estab- lished property of many divalent ions to serve as activators of enzyme systems. The first stage in the metabolic transformation of hexose is the transfer to it of a phosphate ester group from adenosinetriphosphate. This is mediated by the enzyme hexokinase, which has been isolated from yeast. The enzyme is inactive in the absence of magnesium, and a relatively high concentration of this ion is required for full activity, (14). Arginase from both plant and animal tissues is activated by ad- dition of Mn++, Co++, or Fe++; of these M11++ is the most effective (18). The carboxylase of Proteus vulgaris which catalyzes the oxidative decarboxylation of pyruvic acid to acetic acid and carbon dioxide is activated by the addition of Mn++, Mg++, Fe++, Co++, Ni++, and Zn++ (50). Of special interest is the manganese activation of P-car- boxylases since these enzymes catalyze reactions causing assimilation of carbon dioxide and leading to the formation of di- and tricarboxylic acids of importance in intermediary metabolism (46). Daniel I. Anion 329 Space will not permit the mention of many other instances of enzyme activation by metals which have come to light in recent years. Several general conclusions, however, seem apparent. In some cases the activa- tion appears to be specific for one element or at least it is greatly more efficient with it than with any other. This was shown to be the case for magnesium and manganese. Since both of these elements are known to be essential for growth, the evidence of their importance in specific enzyme reactions can be taken as an elucidation of their metabolic function. In instances where several cations activate a given system the evidence does not permit an unequivocal decision as to which element is the actual activating agent in vivo. Nor indeed is there any compelling reason to assume a priori that in the living cell, reactions of this kind must always be catalyzed by one element only. The situation becomes even more complex when work with isolated enzymatic systems brings to light the activating properties of elements such as nickel which are not known to be essential for life. It is possible that for certain cellular reactions specificity of the activator is not a biological requirement. If this view is to be accepted, how can it be reconciled with the well-established specificity of each of the essential elements in growth experiments? It may be assumed that among the several functions which an essential element performs there is at least one for which it is specifically required, no substitution being possible. This view would retain the concept of essentiality in the sense that no sum of partial substitutions for individual reactions, assuming that they all became known, could ever succeed in replacing an essential element. It would, of course, be in accord with the known facts for most of the essential elements. Magnesium, for example, acts as a nonspecific activator of certain enzyme systems though it is also a specific activator for others and no other element can take its place as a component of chlorophyll. This hypothesis would also be compati- ble with reports of "beneficial" effects on the plant from the addition of nonessential elements. Sodium, for instance, could partly substitute for potassium in the sense that it could take over, at least in part, one of its functions. Beneficial effects from adding nonessential elements would thus merely reflect suboptimal conditions with respect to the supply of the essential elements. According to this view a nutrient 330 Mineral Nutrition of Plants medium possessing an optimum supply and favorable conditions for the absorption of the essential elements cannot be improved by the addition of nonessential elements, subject of course, to such reserva- tions about possible unknown micronutrients as were discussed earlier. The discussion thus far has tended to show how, in many instances, the study of function has strengthened or clarified the conclusions derived from growth experiments about the indispensability of inor- ganic nutrients. Recent studies in our laboratory have also shown, however, that conclusions drawn from growth experiments can greatly aid in the interpretation of biochemical observations on function. Be- cause of the pertinence that these results (12) have to the problems under discussion, it is proposed to relate them in some detail. It should be stated at the outset that the objective of the investigation which yielded the results to be examined was somewhat different from the topic under discussion. For the past several years we have been interested in exploring the possible function that inorganic elements, already recognized as essential for plant growth, may have in photo- synthesis. We were encouraged to embark on this investigation by the important discovery of Hill (20) that the long-known capacity of isolated chloroplasts to evolve oxygen can be greatly enhanced by the use of suitable oxidants. Here was a subcellular system which retained the ability to carry out in vitro the photochemical reaction peculiar to the photosynthesis of green plants: the evolution of oxygen resulting from the splitting of water through the capture of the energy of light. The evidence in favor of the identity of the oxygen-liberating mechan- ism in isolated chloroplasts with that in the intact green cells has recently been reviewed by Holt and French (24) and further elaborated by Arnon and Whatley (//). The photochemical evolution of oxygen by chloroplasts isolated from sugar beet and spinach was recently investigated by Warburg and Liittgens (52) who reached the rather striking conclusion that the chloride ion was a coenzyme essential for the photochemical reactions in photosynthesis. That such a simple yet important fact escaped the notice of all the workers in this field was indeed cause enough for Warburg and Liittgens (52) to remark how rash were all the previous theories on the mechanism of photosynthesis. The evidence which led Daniel I. Arnon 331 these authors to conclude that chloride is a coenzyme of photosynthesis was as follows. The isolated chloroplasts lose their capacity for oxygen evolution after several washings in water. They can be reactivated, however, by adding the cytoplasmic fluid. The factor in the cytoplasmic fluid responsible for the reactivation of the chloroplast was found to be heat-stable. An analysis disclosed that the cytoplasmic fluid contained chloride in 0.08 molar concentration. The addition of chloride alone as M/150 potassium chloride brought about complete reactivation. Of the other anions tried, bromide was almost as effective, iodide and nitrate much less so, and fluoride, sulfate, thiocyanate, phosphate, and all the cations tried were without effect. Since chloride was the effec- tive anion found in sufficient concentration in the cytoplasmic fluid, Warburg and Liittgens concluded that it was the natural coenzyme of photosynthesis. Impressive as this chain of biochemical evidence was in support of chloride as a coenzyme of photosynthesis, it posed at once a rather per- plexing physiological problem from the standpoint of plant nutrition. Chloride is not generally regarded as an essential element for the growth of higher plants. Is it then possible that plants can get along in nutrient solutions without a coenzyme required for photosynthesis, a process indispensable for growth ? The fact that Warburg and Liittgens found appreciable amounts of chloride in their plants was not surprising. Chloride is widely distributed in soils and readily absorbed by most plants. Its presence in the plant, however, was hitherto regarded as incidental. We undertook to investigate the problem by growing sugar beet and chard, in nutrient solutions without chloride.* As was expected the plants made excellent growth in the nutrient solution to which no chloride was added. The chloroplasts from these plants were isolated (4) and their oxygen evolution under the influence of light was measured manometrically, by a technique (//) similar to that used by Warburg and Liittgens. Our results disclosed important areas of agreement with those of Warburg and Liittgens as well as several differences. An analysis of *These data have been published separately: D. I. Arnon, and F. R. Whafley, Science, 110:554 (1949). 332 Mineral Nutrition of Plants both the chloroplasts and the cytoplasmic fluid showed no chloride in either, as would be expected in plants grown without chloride. The chloroplasts, even without washing, showed only feeble oxygen evolu- tion. Unlike the experience of Warburg and Liittgens, the addition of the cytoplasmic fluid failed to reactivate them, but it was already noted that our cytoplasmic fluid contained no chloride. On the other hand, we fully substantiated the finding of Warburg and Liittgens that the addition of chloride brought about the activation of chloroplasts, giving us stoichiometric yields of oxygen in relation to the oxidant used. The effect of chloride on the course of oxygen evolution by illuminated chloroplasts is shown in Figure I. We also confirmed the findings of these authors with regard to the influence of other anions on oxygen evolution (Figure i). Bromide had an activating effect about equal with chloride; nitrate and iodide were much less effective; and sulfate, phos- phate, thiocyanate, and acetate were without effect. How should these results be interpreted? The intact plant is able to carry on normal photosynthesis without chloride, as judged by its ex- cellent growth despite the absence of this ion either in the nutrient medium or in the leaf tissue. Yet when chloroplasts are isolated from the same plant, they require chloride for the vigorous progress of the photochemical reaction. One explanation would be that chloride acts in the leaf as a micronutrient and that minute amounts of chloride which would escape detection by the usual chemical analysis are never- theless present in the nutrient medium as an impurity and find their way to the leaf. This explanation, however, although it cannot be ruled out entirely, is rendered unlikely by the data presented in Figure 2. In this chart the rate of oxygen evolution by illuminated chloroplasts (QoJ) 1S plotted against chloride concentration. It will be seen that, whereas small additions of chloride brought about appreciable activa- tion, a fairly high concentration, around 0.007 A/, *s required for full activation. This is in agreement with the value of M/150 potassium chloride reported by Warburg and Liittgens (§2) as necessary for full activation in their experiments. Such relatively high concentrations of chloride are not uncommon in soil-grown plants, but there is strong evidence from these and numerous other experiments that plants can make excellent growth without the presence of measurable amounts of 10 14 Minutes Figure i. Effect of anions on oxygen evolution by illuminated chard chloroplast fragments. Curve I, effect of Cl~; Curve II, effect of Br~; Curve III, effect of N03-; Curve IV, effect of I~; Curve V, NaF; Curve VI, control. Values for sulfate, thiocyanate, and acetate coincided with those for the control. A potassium salt of each re- spective anion was added to give a concentration of io~2M in the manometer vessel, except that the fluoride was added as NaF. Re- action mixture: a chloroplast suspension containing 0.5 mg. of chlor- ophyll, M/15 phosphate buffer, quinone as oxidant. Illumination at flask level approx. 28,000 lux. t = 15° C. Other details of technique were similar to those previously described. — Arnon and Whatley (71). 334 Mineral Nutrition of Plants 1500- 1000 - .chl 500 0.002 0.006 0010 KCI (M) ~yss^- 0.067 Figure 2. Effect of KCI concentration on rate of oxygen evolution by illuminated chloroplast fragments. Q0^1 = cubic millimeters of oxygen, per hour, per milligram of chlorophyll, computed from data obtained for the six-minute period from i min. to 7 min. after turning on the light, t = 200 C. Con- ditions not specified were similar to those given in the legend for Figure i. chloride either in the nutrient medium or in the plant. The other anion capable of giving full activation of photochemical oxygen evolution, bromide, although it is readily absorbed and tolerated by plants in appreciable amounts, is not a common constituent of plants or soils, and there is even less reason for suspecting it as being essential for plant growth. If the view that chloride or bromide is a coenzyme of photosynthesis in vivo is to be abandoned, how can the effect of these anions in vitro be explained? We have formulated the hypothesis that, while in the intact green cell photosynthesis goes on without the participation of either chloride or bromide, once the cell is broken, there is a rapid light-induced deterioration of some cellular substance essential for the photochemical evolution of oxygen by chloroplasts. Chloride or bromide is able to protect this substance against inactivation, but the intact cell accomplishes this in some other manner. This would explain the superfluousness of the halide in the in vivo system as con- trasted with its requirement in the in vitro system. Daniel 1. Anion 335 The hypothesis was tested in the following manner. Isolated chloro- plast fragments were illuminated, without, however, adding the oxi- dant (in this case ferricyanide) which is necessary for the evolution of oxygen to take place. In one instance, chloride was added to the illumi- nated chloroplasts; the control contained no chloride. After twenty minutes of pre-exposure to light, the oxidant was added and the photo- chemical oxygen evolution was measured manometrically. To the ro E E O "O O c\j 20 30 Minutes Figure 3. Protective effect of chloride on illuminated sugar beet chloro- plast fragments. Circles: Illuminated for 20 min. in the presence of chloride. At t = 20 min. tipped in the oxidant (ferricyanide) to the manometer vessel. Crosses: illuminated for 20 min. in the absence of chloride. At t = 20 min. chloride and ferricyanide were added simultaneously. The con- centration of chloride was 0.01M KC1. 1.5 X io-7 moles of K.Fe(CN)6 was added to each vessel. Conditions not specified were the same as those given in the legend for Figure 1. 336 Mineral Nutrition of Plants chloroplast suspension which was exposed to light in the absence of chloride, this anion was added simultaneously with the oxidant. The results are shown in Figure 3. The pre-exposure to light in the absence of chloride inactivated the oxygen evolution system of the chloroplasts. This inactivation was irreversible. The subsequent addition of chloride had only a slight reactivating effect. On the other hand, a vigorous oxygen evolution giving stoichiometric yields resulted from the chloro- plasts which had received added chloride during their exposure to light. Thus, chloride appeared to exert protective action on some essen- tial photosynthetic factor that in the absence of this anion was irreversi- bly destroyed by light. Chloride also seemed to exert some protective action on the chloroplasts in the dark. There was evidence of inactiva- tion from shaking the chloroplasts in the manometer vessels at 15 ° C. for a period equal to the light exposure. The inactivation in light, how- ever, was much more pronounced. It goes without saying that the identification of this substance would be of great physiological interest. Experiments along this line have been under way in our laboratory, but no statement is possible at this time. In addition to chloride, Warburg and Liittgens (52) reached a con- clusion of great significance with regard to zinc. They considered that the photochemical evolution of oxygen by chloroplasts is catalyzed by a metal and that in all probability the metal concerned is zinc. The evidence for this conclusion was as follows. The oxygen evolution re- action was strongly inhibited by o-phenanthroline, a well-known metal complex former; this reagent forms complexes with bivalent iron, nickel, cobalt, and zinc. The o-phenanthroline inhibition was reversed and the chloroplast fully reactivated by adding an excess of zinc. (In the example cited by Warburg and Liittgens, ten times as much as would be required to bind the o-phenanthroline.) The addition of iron brought only partial reactivation. If, however, the metal was added to o-phenanthroline prior to placing it in contact with the chloroplasts, then zinc was not the only element which prevented inhibition: iron was equally effective, as were divalent cobalt and nickel ions. Warburg and Liittgens (52) reasoned that since o-phenanthroline inhibition was observed only when this reagent was still free to combine with the metals in the chloroplasts, it is probable that some metal is involved as Daniel I. Anion 337 a catalyst in the oxygen evolution. They analyzed for the inorganic constituents of chloroplasts including iron and zinc and, on the basis of the amount of zinc found and the amount of inhibitor added, con- cluded that zinc was the metal concerned in the oxygen evolution re- action of chloroplasts. The problem of o-phenanthroline inhibition of the oxygen-evolution reaction has been independently investigated in our laboratory as part of a general study of chloroplast reactions (12). In agreement with Warburg and Liittgens, we found that o-phenanthroline inhibition was fully reversed by an excess of zinc and much less effectively by iron. Zinc was also very effective in reversing the inhibition even when added not in a tenfold excess but in a stoichiometric amount in relation to o-phenanthroline (Figure 4). However, when added in stoichiometric amounts, two other metals not known to be essential for plant growth, nickel and cobalt, were found to be even more effective in reversing o-phenanthroline inhibition. In our experience these two metals not only protected by binding the o-phenanthroline before it was mixed with the chloroplasts as observed by Warburg and Liittgens (52), but also reactivated the previously inhibited preparations in a manner simi- lar to zinc (Figure 4). Reactivation was also obtained with copper, a micronutrient of estab- lished status, known to occur in chloroplasts. The reactivating efficiency of copper was, on the basis of stoichiometric amounts, intermediate be- tween iron and zinc. Unlike iron,* however, an increase in the concen- tration of copper gave total reactivation in a manner similar to zinc. As seen in Figure 4, doubling the concentration of copper brought about a marked increase in its effectiveness. It was found then that the reversal of o-phenanthroline inhibition was accomplished by two elements, nickel and cobalt, not known to be essential for plant life and by two, copper and zinc, recognized as micronutrients. It is considered unwarranted, on the basis of evidence now available, to associate any one metal with the photochemical evo- lution of oxygen. Deductions on the basis of composition of plant tis- sues appear to be, in the light of previous discussion, not wholly reliable. *After this manuscript was submitted for publication, new evidence on this point became available. It will be published elsewhere. to . E E Q> O 3 O Cl CM 8 12 Minutes Figure 4. Effect of metals on reversal of o-phenanthroline (o-P) inhibi- tion of sugar beet chloroplasts. o-Phenanthroline was added to chloroplast fragments in all cases except to the control represented by Curve I. Curve II, effect of adding nickel and cobalt and doubling the copper concentra- tion; Curve III, effect of zinc; Curve IV, effect of copper; Curve V, effect of iron (ferrous); Curve VI, o-phenanthroline alone. The addition of chromate was without effect. The concentration of o-P was 5 X io~5M. The metals were added as sulfates and nitrates in a 5/3 X io— 5M con- centration to give a stoichiometric ratio of 3 o-P: 1 metal. Conditions not specified were the same as those given in legend for Figure 1. Daniel I. Anion 339 This it not to say that zinc is not the metal specifically concerned in the photochemical reactions of photosynthesis, but to suggest that such a conclusion is as yet not supported by incontrovertible evidence. It is hoped that experiments now in progress may throw some more light on the subject. The foregoing discussion has attempted to demonstrate that an inte- grated concept of the essentiality of mineral elements rests on the com- bined contributions from studies of inorganic requirements for growth and investigations on functional aspects of the essential nutrients. The two approaches, which for the sake of convenience may be designated as the physiological and biochemical, respectively, are mutually supple- mentary. It was shown how in the case of chloride, its proposed essen- tial status as a coenzyme for photosynthesis could not be accepted in the light of evidence from growth experiments. It is possible, however, that in the future an insight into the essential function of an inorganic element may be gained from biochemical studies, well in advance of any knowledge gained from growth experiments. A case in point is cobalt in animal nutrition. Cobalt was recognized as a micronutrient essential for ruminants (j6), but there was no evidence that it was re- quired by other animals. Recently a striking discovery was made that cobalt is a component of vitamin B12, which is identified with the "anti- pernicious anemia" and the "animal-protein" factors (42,43,41). Vita- min B12 seems to be essential for all animals, and cobalt, which was not previously found in a compound of a natural origin, must therefore now also be regarded as an essential micronutrient for animals other than ruminants. (However, recent experiments indicate that in rumi- nants at least, cobalt may have functions distinct from its association with vitamin B12; Becker et al„ /j). Evidently the quantitative re- quirement for cobalt by nonruminants is so small that it escaped de- tection in direct growth experiments. An idea as to the experimental difficulties which may be involved is given by the fact that four tons of liver were required to yield one gram of the pure vitamin (42). The study of the inorganic requirements of higher plants has, apart from its richly rewarding scientific purpose in contributing to the un- derstanding of plant growth and metabolism, an important practical objective as well. Developments in this field have provided in the past, 340 Mineral Nutrition of Plants and will continue to provide in the future, a scientific basis for fertili- zation practices. As the list of essential elements has expanded, it was possible to answer with increasing assurance how many indispensable nutrients need to be supplied in the external medium to insure optimal plant growth. It is hoped that advances in our knowledge of the func- tion of inorganic nutrients will provide a sound basis for estimating quantitative requirements of fertilizer elements for different crops at different stages of growth and in relation to climatic factors. REFERENCES i. Arnon, D. I., Soil Sci., 44:91 (1937). 2. , Am. }. Botany, 25:322 (1938). 3. , Science, 92:264 (1940). 4. , Plant Physiol., 24:1 (1949). 5. , and Hoagland, D. R., Science, 89:512 (1939). 6. , and Hoagland, D. R., Soil Sci., 50:463 (1940). 7. , and Meagher, W. R., Soil Sci., 64:213 (1947). 8. , Simms, Helen D., and Morgan, Agnes Fay, Soil Sci., 63:129 (i947)- 9. , and Stout, P. R., Plant Physiol., 14:371 (1939). 10. , and Stout, P. R., Plant Physiol., 14:599 (1939)- 11. , and Whatley, F. R., Arch. Biochem., 23:141 (1949). 12. , and Whatley, F. R., Unpublished data (1949). 13. Becker, D. E., Smith, S. E., and Loosli, J. K., Science, 110:71 (1949). 14. Berger, L., Stein, M. W., Colowick, S. P., and Cori, C. F., /. Gen. Physiol., 29:379 (1946). 15. Bertrand, G., Compt. rend., 124:1032, 1355 (1897). 16. Bradfield, J. R. G., Nature, 159:467 (1947). 17. Chandler, W. H., Botan. Gaz., 98:625 (1937). 18. Folley, S. J., and Greenbaum, A. L., Biochem. ]., 43:537 (1948). 19. Hewitt, E. }., and Jones, E. W., /. Pomol. Hort. Sci., 23:254 (1947). 20. Hill, R., Proc. Roy. Soc. (London), 1276:192 (1939). 21. Hoagland, D. R., Proc. Am. Soc. Hort. Sci., 38:8 (1941). 22. , Lectures on the inorganic nutrition of plants (Chronica Bo- tanica, 1944). 23. , and Snyder, N. C, Proc. Am. Soc. Hort. Sci., 30:288 (1933). 24. Holt, A. S., and French, C. S., in J. Franck and W. E. Loomis, Editors, Photosynthesis in Plants (Ames, Iowa: Iowa State College Press, 1949), chapter 14. 25. Horner, K. C, Burk, D., Allison, F. E., and Sherman, M. S., /. Agr. Research, 65:173 (1942). Daniel I. Anion 341 26. Javillier, M., Compt. rend., 155:1551 (1912); 158:140 (1914). 27. Jensen, H. L., Proc. Linnean Soc. N. S. Wales, 72:73 (1947). 28. Keilin, D., Proc. Roy. Soc. (London), 698:312 (1925). 29. , and Hartree, E. F., Proc. Roy. Soc. (London), 6127:167 0939)- 30. , and Mann, T., Proc. Roy. Soc. (London), 6125:187 (1938). 31. ■, and Mann, T., Nature, 143:23 (1939). 32. , and Mann, T., Nature, 144:442 (1939). 33. Kubowitz, F., Biochem. Z., 299:32 (1938). 34. Lepierre, C, Compt. rend., 156:258 (1913). 35. McHargue, J. S., /. Am. Chem. Soc, 44:1592 (1922). 36. Marston, H. R., Ann. Rev. Biochem., 8:557 (1939). 37. Maze, P., Compt. rend., 160:211 (1915). 38. Mulder, E. G., Plant and Soil, 1:94 (1948). 39. Piper, C. S., Austr. Inst. Agr. Sci., 6:162 (1940). 40. Raulin, J., Ann. Sci. Nat. Bot., 5 Ser., 11:93 (1869). 41. Rickes, E. L., Brink, N. G., Konuszy, F. R., Wood, T. R., and Folkers, K., Science, 108:134 (1948). 42. Smith, E. L., Nature, 161:638 (1948). 43. •, Nature, 162:144 (1948). 44. Sommer, Anna L., Plant Physiol., 6:339 (1931). 45. , and Lipman, C. B., Plant Physiol., 1:231 (1926). 46. Speck, J. F., /. Biol. Chem., 178:315 (1949). 47. Steinberg, R. A., Am. J. Botany, 6:330 (1919). 48. Stiles, W., Trace elements in plants and animals (Macmillan, 1946). 49. Stout, P. R., and Arnon, D. I., Am. J. Botany, 26:144 (1939). 50. Stumpf, P. K., /. Biol. Chem., 159:529 (1945). 51. Tissieres, A., Nature, 162:340 (1948). 52. Warburg, O., and Luttgens, W., Biochimia (U.S.S.R.), 11:303 (1946). 53. Warrington, K., Ann. Botany, 37:629 (1923). CHAPTER I *r Mineral Nutrition in Re- lation to the Ontogeny of Plants W. F. LOEHWING T he study of mineral nutrition in relation to ontogeny is merely one type of approach to an understanding and control of plants as organisms. A study of this sort becomes primarily an attempt to arrange the known facts of mineral metabolism as a progressive se- quence of events as these relate to the commonly observed growth pat- terns of plants. The major problem is one of integration of somewhat isolated data with the growth process as a whole. It seems desirable to outline the accumulation and distribution of inorganic ions and then to trace as far as possible their connection with tissue differentiation and the progression of the entire life cycle of the plant as a whole. An understanding of mineral nutrition in relation to ontogeny is complicated by the fact that numerous meristems give rise continu- ously to new tissues and organs, with the result that various parts of the plant are usually in different stages of development. Thus, com- parable size and age of parts or of the plant as a whole are not neces- sarily criteria of developmental similarity (121). At mid-maturity of a plant, its various parts may display the entire range of development from youth to senescence (//). Upon these developmental differences between parts are superimposed certain ecological contrasts (9^). During growth, a single plant often modifies its environment in such a way that organs appearing in a chronological sequence commonly form an ecological succession (76, 152). Due to the progressive modification of apical meristems in the course of vegetative development, for example, there usually occurs an axial progression from mesomorphy of the 344 Mineral Nutrition of Plants lower leaves to comparative xeromorphy of upper leaves on the same stem (y6, 84). It is also obvious that data on root physiology are essen- tial to a comprehensive grasp of ontogeny. This fact explains the prefer- ence of investigators for liquid and gravel cultures from which roots are readily retrievable. Such cultures, in addition, permit better control and study of the substrate than in the case of soil-grown plants. Finally, a running inventory on the composition and behavior of the nutrient solution or substrate is a great aid to an understanding of the plant's relationship to and effect upon its edaphic environment (/, 14,26, jy, $y, 59,60,106). Thus, Hartel (52) reports translocation of carbon dioxide from roots to shoots where it can be used in photosynthesis, and Ulrich (7^5) has noted the buffer action of organic acids in roots when cation absorption exceeds anion intake (79). In scanning the numerous studies of developmental physiology, it is surprising to find many similarities in fundamental processes despite the extreme diversity of plant types and environmental conditions in- volved. As might be expected, a preponderance of data exists concern- ing herbaceous annuals, and the most consistent portion of this infor- mation relates to plants grown under controlled environmental con- ditions. Comparable data for biennials and perennials, often of neces- sity grown under field conditions, are more difficult to obtain, hence, often relatively incomplete and more difficult to correlate and interpret. Because of this fact the present discussion is limited to those aspects of developmental processes which seem common to the wide range of ordi- nary herbaceous annuals most frequently investigated. THE VEGETATIVE STAGE Under favorable growing conditions, the phase of rapid vegetative enlargement in typical annuals is characterized by progressive incre- ments in absolute amounts of inorganic elements, carbohydrates, and proteins. Due to the accelerated synthesis of organic compounds, the proportion of ash on a percentage basis begins to fall even though abso- lute amounts of the latter may continue to rise until well toward ma- turity (29, 7/, log). Many annuals tend to absorb the major portion of their total mineral supply in very early life (iy.yj, 142) and early ab- sorption is, in general, in excess of current needs when the external W.F.Loehwing 345 supply is favorable. In early vegetative stages under conditions of bal- anced or constant supply, nitrogen, potassium, and phosphorus com- monly increase faster than calcium, iron, magnesium, and sulfur due in part to the relative immobility of the latter elements within the plant (3, 24, yi, 111 , 112). It may be noted, however, that the rate of absorp- tion of a particular ion is determined not only by its availability in the substrate but also by the concentration thereof already in the plant (20, 47, ji, 136, 75/). Numerous investigations (//, 75) indicate that the beneficial effects of soil fertilization are due primarily to increase in foliar area and of assimilative tissue rather than to increase in effi- ciency of assimilative processes. As might be expected in early stages of development when new tis- sues are being formed on a large scale, the young plant is relatively more active in accumulation of water (77, 55, 68, 99, 727, 140), nitrogen, and protein than in its later life (77, 54, 62, y^, 112, 138). On the basis of large amounts of inorganic and amide nitrogen found in young plants under favorable conditions, much of the early intake is in the nature of luxury consumption (47). Experimentally, this is made evi- dent by the fact that gradients of curves for nitrogen accumulation are usually steeper than those for dry weight gains in young plants (77, 6y, j$, 112, 138). Recent work on growth substances and naturally occurr- ing phytohormones indicates important interrelationships of such sub- stances with nitrogen (9, 10), zinc (726, 134), and several other nutrient elements (27, 81). Auxin activity diminishes under conditions of nitro- gen deficiency before lack of nitrogen causes retardation of growth or stem elongation {133)- It has been observed that zinc controls trypto- phane formation and, hence, auxin activity because tryptophane is an auxin precursor. Went (141) discusses the work of Bonner and others on the vitamin requirements of flax, pea, and tomato roots. Flax roots which require pyrimidine and thiazole for growth do not themselves synthesize these thiamin precursors but are normally dependent upon translocation of them from the shoot which can produce them. Root tissue-cultures of flax grow only when pyrimidine and thiazole are added. In this instance, sulfur functions as an organic complex in thia- min. Similar relationships seem to prevail in pea and tomato roots for nicotinic acid and vitamin B6, respectively {141). These observations 346 Mineral Nutrition of Plants raise the question whether failure of nutrient absorption by roots when their carbohydrate supply is low, may not also be associated with failure of hormone translocation to roots from the shoot. During the vegetative phase of growth, there is an intimate connec- tion between carbohydrate and protein synthesis. Not only are carbo- hydrates and nitrogen used in the synthesis of proteins, but a portion of the soluble hexoses evidently provides the respiratory energy neces- sary for the chemical reduction of nitrates as an antecedent to amino acid and protein formation. In fact in young plants, the supply of solu- ble sugars appears largely to condition the rate of protein synthesis (85, 86, 8y, 12J, 753). Obviously, the availability of oxygen also is essen- tial to the respiratory oxidation of a portion of the carbohydrates but this usually is not a limiting factor in early growth as it is later. Mothes (85, 87) has shown that all conditions, such as light, photosynthesis, and open stomata, which tend to raise internal oxygen tension, favor protein formation. Conversely, protein hydrolysis is accelerated by a low in- ternal oxygen tension and by low water content in later development (25). Thus, the rate of photosynthesis as a source of both carbohydrates and oxygen is closely bound up with nitrate reduction and protein syn- thesis. During the later phases of active vegetative growth, the plant rapidly accumulates carbohydrates and appears to become relatively less efficient in protein than in carbohydrate elaboration (//) as shown by increments in the carbon-nitrogen ratio (55). Thus from the period of germination to flowering, three fairly dis- tinct stages of metabolism are evident. The conspicuous stages in the nutrition of the vegetative plant comprise an initial anabolic phase (I) in which intake of inorganic nutrients and synthesis of proteins is rapid. In the second phase (II), the accumulation of carbohydrates accelerates while the rate of protein synthesis gradually diminishes. As flowering is approached, a third or catabolic phase (III) becomes evident in which hydrolysis of reserves begins to overbalance synthesis and a general internal redistribution of nutrients is initiated. Though conditions of environment and nutrient supply determine to a considerable degree the exact time of the shift from predominantly anabolic to catabolic activity, the latter transition is characteristically associated with flower- ing and commonly initiated prior to anthesis (40). W. F. Loehwing 347 THE FLOWERING STAGE In reference to the initiation of reproductive processes, recent data (/j, 18, 80, 96, 144, 145. 150) have confirmed and elaborated many older observations to the effect that the phenomenon of synapsis or sporo- genesis represents a turning point in nutritional metabolism (55, 56, 61, 65, 141). (a) In monoecious species such as corn, it has been observed that the origin of staminate and pistillate organs is associated with a transitory but systemic acceleration of anabolic over catabolic processes including more rapid salt absorption and dry weight gains. (/?) Under normal conditions the metabolic stimulus associated with synapsis is brief and soon gives way to a reduction in anabolic processes during the ensuing phase of blossoming or anthesis. The flowering phase is usually characterized by a subsidence of anabolic activity as well as the inaugu- ration of fundamental modifications (jo) and redistribution of organic and inorganic nutrient components (11, 25, 40, 61, y8). Subsequent events vary considerably among species, some of which, for example, undergo no further elongation of the main axis following anthesis (//). While there are important differences between plants of determinate habit, such as grasses, and those of indeterminate habit, as in many dicots, the subsidence in rate of stem elongation can, nevertheless, often be taken as an index of the fact that reproduction is already under way even though gross morphological evidence of floral parts is not yet visible (j], 104). (c) The decline of anabolic activity associated with blossoming gradually gives way to what is usually the final resurgence of absorption of mineral nutrients and acceleration of organic synthesis in vegetative tissues. This anabolic stimulus is associated with the fusion of male and female nuclei in syngamy and the very early enlargement of young fruits (22, 38, 95. 96, 144, 145. /50). From the standpoint of ontogeny, much could be said in favor of beginning such a discussion with sexual fertilization or syngamy which, after all, is the actual inception of the new plant. Such a procedure would be justified not only by the chronological sequence of events but by the fact that many environmental factors to which the growing embryo is exposed prior to its development into a mature and dormant seed can predetermine in considerable degree the course of ontogeny subsequent to seed germination. 348 Mineral Nutrition of Plants As stated, the onset of blossoming or anthesis is marked by an appar- ently simultaneous reduction in absorption by roots, an internal shift in water balance (//, 37, 40, 56, 58, 63, 64, 107, 121, 122, 127), and redis- tribution of both organic and inorganic nutrients (28, 73, 89, (jo, 127). At the present time it seems that the foregoing phenomena are all the results of some as yet indefinitely defined regulatory or causative fac- tors. We already have some evidence, however, that they are at least associated with, if not caused by, increments in growth substances at reproductive loci within the plant following syngamy {144, 145, 150). During anthesis, the carbohydrate supply of roots and their rates of absorption commonly fall to low levels (//, 25, 40, 4], 61, 78, 102, 137). The inadequacy of root carbohydrates at this period has frequently been advanced as the cause of their low absorptive activity (40). As already noted, however, recent data indicate that the organic reserves of the root function jointly with growth substances in absorption and root enlargement. In liquid and gravel culture experiments in which entire plants in- cluding roots and the nutrient solution have been analyzed, there is evidence of transitory excretion of certain mineral elements during anthesis {4, 22, 38, 47, 5/, 147). Nitrogen and potassium especially often increase temporarily in the nutrient medium (5/). The recent work of Cailachian (23) on nitrogen in relation to flowering indicates that there are three categories of plants — namely, those in which nitrogen acceler- ates, delays, or has no effect on flowering (144). During anthesis, lower leaves show a rapid acceleration in loss of water (82, 83, 121) and organic reserves (83), a trend which subse- quently reaches the extreme of protoplasmic disintegration and transfer of such cellular residues to reproductive organs and to younger vege- tative tissues. In the early stages of this trend, loss of water is progres- sive despite appreciable increments in osmotic values of press sap. Hydrolysis of insoluble organic reserves of lower leaves is a conspicuous phenomenon at this time (2), yet the resulting rise in the osmotic pres- sure of tissue fluids is usually unable to arrest water loss (//, 83, 103, 127). The inference follows that factors other than osmotic pressures serve in regulation of the water balance of tissues (83, 127). Numerous studies on the drought resistance of plants have shown their extreme W. F. Loe hiving 349 sensitivity to water shortages during flowering {2J, 68, 116, i2g). Even if plants survive drought at the flowering stage, they seldom show a normal course of development thereafter. Smirnov {127) and his associates have shown by a series of ingenious experiments that water loss and hydrolysis of organic reserves occur simultaneously with a reduction in the amount of hydrophilic colloids in vegetative tissues. Smirnov stresses the "salting out" effect of mineral nutrients upon organic colloids and the resultant reduction in water retentivity by colloids after precipitation. Precipitation of cell colloids is evidently accompanied by liberation of previously absorbed enzymes and a simultaneous increase in their hydrolytic action. During this period, there also occur marked changes in pH of tissue fluids; such alterations may be additive to the effects of colloid precipitation in in- creasing the hydrolytic action of enzymes. Many investigations provide evidence of marked reduction in the rate of photosynthetic activity during the period of rapid hydrolysis of organic reserves (//, 148). Relatively low levels of tissue moisture and hydrolysis of organic re- serves in lower leaves frequently involve living protoplasm (46, 49, 99, 108, 123, 124), and, once the protoplasmic components of such tissues begin to undergo hydrolysis, they usually become incapable of renewed synthesis even upon amelioration of the water and nutrient supply (Sj, 104, 132). This observation is usually offered as the explanation of the early death and abscission of lower leaves of plants in the repro- ductive phase. Mothes (8y) has shown the dependence of protein syn- thesis upon internal oxygen tension, and a reduction in rate of protein formation commensurate with the decline in the rate of photosynthesis. Smirnov's data {i2j), in turn, show a correlation between the rates of respiration and protein synthesis (49). It thus appears that photosyn- thetic oxygen favors protein formation by acceleration of aerobic oxi- dation of carbohydrates. The resultant energy is important to the re- duction of nitrates as well as to the union of nitrogen with carbo- hydrate derivatives in amino acid synthesis. Smirnov points out that in the early life of annuals, protein synthesis parallels, and hence is pre- sumably dependent upon, the concentration of hexose sugars. With the onset of reproduction, however, protein production ceases to be propor- tional to the soluble sugar content but parallels the rate of foliar res- 350 Mineral Nutrition of Plants piration. Thus, in the vegetative phases, carbohydrate supply is the con- trolling factor in protein synthesis, while oxygen supply becomes a major regulatory factor in the synthesis of proteins during the repro- ductive phase. In addition to the foregoing functional factors influencing the water economy of the flowering shoot, vascular tissues also undergo compli- cating structural modifications (30, 31, 74, 113, 125, 130, 131, 143, 149). The work of several investigators (116, ijq) reveals a subsidence in cambial activity which begins in the vicinity of floral buds and com- monly extends progressively toward the base of the stem (125). Phloem formation especially seems to be reduced and failure of vascular differ- entiation commonly involves pedicels or fruits stalks, thus often im- pairing fruit setting or fruit enlargement (27). The impairment of con- duction appears during the flowering stage and retards the redistribu- tion of nutrients; the reduction in conductive capacity of vascular bun- dles in stems and pedicels may become a temporarily limiting factor in the rate of growth of the shoot apex and of fruits. THE FRUITING STAGE The fruiting stage has its origin in syngamy. The early stages of fruit enlargement are commonly associated with marked increments in ab- sorption by roots and accelerated anabolism of the younger parts of the shoot (13, 18, 27, 34, 42, 6$, 74, gi, 94, 100, 127, 144, 145, 150). Gains in nitrogen and potassium become appreciably higher (74). Absorption of phosphorus and iron, though fairly steady at first, tend to rise, some- times to surprisingly high levels as maturation supervenes (44, 1 17, 139). There is an increased accumulation of proteins and carbohydrates, the latter usually being the greater in terms of dry weight gains (127). The origin of the systemic stimulus to accelerated activity following syngamy appears to be associated with increments in growth substances at reproductive loci and their translocation to adjacent tissues (8, 88). As would be expected, increasing amounts of organic and inorganic reserves are diverted from vegetative to reproductive organs as more fruits are set and their enlargement accelerates (28, 73, go, 145). It is interesting to observe, however, the comparatively uniform composi- tion of seeds and fruits in relation to the frequently great differences in W.F.Loehwing 351 mineral elements in the vegetative organs nourishing them (5, 105). Seeds and fruits are highly selective in the elements which they accumu- late from leaves and stems (5). In recent years, agronomists studying the problems of fertilizer place- ment have reported significant differences in response to a given ele- ment with the stage of plant development at which it is supplied (6, 7, 12, 16, 19, 32, 33, 39, 41, 45, 48, 50, 72, 97, 101, 105, no, 114, 115, 118, 119, 120, 128). Striking differential effects upon specific tissues and the course of ontogeny have been observed. There is evidence, for example, that nitrogen applications at or immediately following anthesis induce re- sponses in reproductive and vegetative organs which are quite different from those observed in plants held continuously at uniform levels of nitrogen supply (12, 33, 47, 101, no, 114, 128). During the early fruiting stage, many species exhibit a high absorptive capacity for and extreme sensitivity to increments or diminutions in nutrient supply (34, 92, 114). Their responses to fertilizers made available at this time are often wholly unlike those supplied at earlier or later stages (132). Where marked deviations from ordinary growth patterns are observed following changes in nutrient level at mid-developmental stages, it be- comes of interest to learn which tissues are primarily affected and in what manner, both as to function and structure. Detailed information on these responses is as yet quite meager, but judging from definite responses thus far reported, closer study of them should prove peculiarly productive in extending our understanding and control of fundamental features of plant development. The recent work of Rankin (no) on staggered and late supplies of nitrogen in producing differential effects upon the number of spikes, florets per spike, weight and number of kernels per plant in wheat is a case in point (12, 115, 120, 128). Another is the work of Sybil (132) on tobacco in which a shift from low or medium to high nitrogen supply at anthesis produced a leaf structure and organization quite unlike those observed in plants grown at uniform nitrogen levels. It is also worth noting that the form in which nitrogen is supplied is important, as the effects of ammoniacal and nitrate nitrogen may be quite different (154). The yield of fruit produced by a given amount of vegetative tissue can be varied appreciably depending on the form of available 352 Mineral Nutrition of Plants nitrogen and aeration of the substrate (146). The effects of elements other than nitrogen also appear to vary significantly with the stage of development at which applied (101, 105, 144). The phenomena of nutrient balance between vegetative organs and fruits, especially in heavily fruiting varieties such as cotton and tomato, have been described by many investigators. The effects of a heavy crop of fruit in depleting the nitrogen reserves of leaves and stems are rather well known, as also are the cyclic renewals of vegetative activity when fruit abscission or maturation occurs (28, 73, go, 137, 145). The usual course of events in numerous annuals is the depletion of leaves to the point of protoplasmic disintegration and the rapid initiation of senescence terminating in abscission or death of leaves and excretion of mineral elements by roots to substrate (35, 36, 37, 44, 61, 69). In conclusion it may be noted that, though mineral nutrients are not the initial and primary causes of tissue differentiation or inception of reproductive processes, they are often controlling factors in the imple- mentation of the plants' developmental potentialities. Recent work dis- closes the intimate relation of mineral elements to the formative action of growth substances (66), to the respiratory action of nucleotides, and to other intermediates in oxidative metabolism. Thus, the action of phytohormones and other specifically morphogenetic compounds is in- timately correlated with mineral nutrients much as the inorganic ions are related to photosynthesis, enzyme action, and other vital functions in the regulation of growth and the ontogenetic cycle. REFERENCES 1. Achromeiko, A. I., Jahrb. Pflanzenernlihr-Diing. und Boden\unde, 42:156 (1936). 2. Ali-Zade, M., Doklady A\ad. Nau\ S.S.S.R., 30:254 (1941); Chem. Abstracts, 35:8016 (1941). 3. Anderson, P. J., Swanback, T. R., and Street, O. E., Conn. Agr. Expt. Sta. Bull., 422:6 (1939). 4. Andre, G., Compt. rend., 154:1817 (1912); 151:1378 (1910); 156:1164 (1913); 152:965 (1911). 5. Arnon, D. I., and Hoacland, D. R., Botan. Gaz., 104:576 (1943). 6. Avdonin, N. S., Soc. zern. Hoz., 4:127 (1939); Herb. Abstracts, 10: 148 (1940). W.F.Loe •hiving 353 7. , Do^lady Vsesoyuz. A\ad., 11:26 (1939); Herb. Abstracts, 10:148 (1940); Chem. Abstracts, 35:6287 (1941). 8. Avery, G. S., Jr., Berger, J., and Shalucha, B., Am. }. Botany, 29: 765 (1942); Chem. Abstracts, 36:4861 (1942); Chem. Abstracts, 37: 1156 (1943). 9. Avery, G. S., Jr., Burkholder, P. R., and Creichton, H. B., Am. ]. Botany, 24:553 (1937). 10. Avery, George S., Jr., and Pottorf, Louise, Am. J. Botany, 32:666 (i945)- 11. Bakhuyzen, H. L. van de Sande, Food Research Institute (Stanford University, 1937) 4°°- 12. Bereznitskaya, N. I., Compt. rend. acad. sci. U.R.S.S., 30:186 (1941); Chem. Abstracts, 35:6047 (1941). 13. Biddulph, Orlin, and Brown, D. H., Am. J. Botany, 32:182 (1945). 14. Bizzell, J. A., /. Am. Soc. Agron., 10:97 (I91^)- 15. Blackman, G. E., and Rutter, A. J., Ann. Botany, N. S., 12:1 (1948). 16. Boichenko, E., Chemisation Socialistic Agr. (U.S.S.R.), 12:55 (1939); Khim. Referat. Zhur., 6:57 (1940); Chem. Abstracts, 36:5941 (1942). 17. Boresch, K., in Honcamp, F., Handbuch der Pflanzenerndhrung und Diingerlehre (Berlin, J. Springer, 1931) 180. 18. Briggs, G. E., Kidd, F., and West, C, Ann. Appl. Biol., 7:103, 202, 403 (1920); New Phytologist, 19:200 (1920). 19. Brune, F., Boden\imde u. PfJanzenerniihr., 24:1 (1941); Chem. Zentr., 2:1189 (1941); Chem. Abstracts, 38:2155 (1944). 20. Bruns, Walter, /. Landw., 83:285 (1935); Chem. Abstracts, 30:2307. 21. Brunstetter, B. C, Myers, A. T., Mitchell, J. W., Stewart, W. S., and Kaufman, M. W., Botan. Gaz., 109:268 (1948). 22. Burd, J. S., /. Agr. Research, 18:51 (1919). 23. Cailachian, M. H., Compt. rend. acad. sci. U.R.S.S., 40:146 (1945). 24. Campbell, E. G., Botan. Gaz., 78:103 (1924). 25. Chibnall, A. C., Protein Metabolism in the Plant (Yale Univ. Press, I939)- 26. Clements, F. E., Carnegie Inst. Wash. Publ. 315 (1921). 27. Cochran, H. L., Proc. Am. Soc. Hort. Sci., 29:434 (1932); Cornell Univ. Agr. Expt. Sta. Mem., 190:39 (1936). 28. Crowther, F., Ann. Botany, 48:877 (1934); Ann. Botany, N. S., 5: 509 (1941). 29. Czapek, F., Biochemie der Pflanze (Vol. II, Jena, Gustav Fischer, 1905). 30. Dastur, R. H., Ann. Botany, 38:779 (1924); 39:762 (1925). 31. , Ann. Botany, 39:769 (1925). 32. Davidson, J., /. Am. Soc. Agron., 14:4 (1922); 9:145 (1917); 10:193 (1918); /. Agr. Research, 23:55 (1923). 354 Mineral Nutrition of Plants 33. Davis, M. B., and Hill, H., Sci. Agr., 8:681 (1928); 14:411 (1934). 34. Dearborn, R. B., Cornell Agr. Expt. Sta. Mem., 192:26 (1936). 35. Deleano, N. T., Bui. soc. stiinte Cluj, 7:45 (1932); Chem. Abstracts, 27:3501 (1933). 36. , and Andreesco, M. I., Beitr. Biol. Pflanzen, 19:249 (1932). yj. , and Bordeianu, C, Beitr. Biol. Pflanzen, 20:179 (1932); 19: 249 (!932); and 24:19 (1936). 38. , and Vladescu, I. D., Bull. soc. chim. biol., 19:1366 (1937). 39. Demidenko, T. T., and Popov, V. P., Compt. rend. acad. sci. U.R.S.S., I7:59 (1937); ^-i67 0938). 40. Dennison, R. A., Plant Physiol., 20:183 (T945)- 41. Dobrounof, L. G., Compt. rend. acad. sci. U.R.S.S., 19:215 (1938). 42. Doneen, L. D., Wash. Agr. Expt. Sta. Bull. 2g6 (1934). 43. Ergle, D. R., /. Am. Soc. Agron., 28:775 (1936). 44. Fagan, T. W., and Watkin, J. E., Welsh J. Agr., 7:229 (1931). 45. Golle, V. P., and Demidenko, T. T., Compt. rend. acad. sci. U.R.S.S., 27:284 (1940); Chem. Abstracts, 34:7:544 (1940). 46. Golovina, A. S., Voronezh. Gosudarst. Univ., Nauch. Raboty Studen- tov, 51 (1939): Khim. Referat. Zhur., 6:44 ( 1940); Chem. Abstracts, 36:4546 (1942). 47. Goodall, D. W., and Gregory, F. G., Imperial Bureau of Horticul- ture and Plantation Crops, Technical Communication, 17:1 (1947). 48. Gericke, W. F., Botan Gaz., 80:410 (1925); See also Soil Sci., 29:207 (1930); Science, 59:321 (1924); Science, 60:297 (x924)- 49. Gregory, F. G., Ann. Rev. Biochem., 6:557 ( 1937). 50. Grizzard, A. L., Davies, H. R., and Kangas, L. R., /. Am. Soc. Agron., 34:327 (1942). 51. Hall, Wayne C, Plant Physiol., 24:755 (1949). 52. Hartel, Otto, Jahrb. wiss. Botan., 87:173 (1938); Chem. Abstracts, 35:8017 (1941). 53. Hicks, Phyllis A., New Phytologist, 27:1, 108 (1928). 54. Honl, V., Biochem. Z., 225:94 (1930). 55. Hornberger, R., Landw. Jahrb., 11:359 (1882). 56. , Landw. Vers.-Sta., 31:415 (1885). 57. Huber, B., Jahrb. wiss. Botan., 64:1 (1924). 58. Hurd-Karrer, A. M., and Taylor, }. W., Plant Physiol., 4:393 (1929). 59. Jakobey, I., Kiserletiigyi K6zu°menye\, 43:13 (1940); Chem. Ab- stracts, 35:5152 (1941). 60. Jenny, H., Overstreet, R., and Ayers, A. D., Soil Sci., 48:9 (1939). 61. Knowles, F., and Watkin, J. E., /. Agr. Sci., 21:612 (1931). 62. Knowles, Frank, and Watkin, }. E., /. Agr. Sci., 22:755 (1932). 63. Konig, J., Hammerbacher, Fr., and Brimmer, C., Landw. Jahrb., 5:657 (1876). W.F.Loehwing 355 64. Kreusler, U., Prehn, A., and Hornberger, R., Landw. ]ahrb., 8:617 (1879). 65. , Prehn, A., Hornberger, R., and Becker, G., Landw. Jahrb., 6:759 (1877); 6:787 (1877); 7:536 (1878); 8:617 (1879). 66. Leopold, A. C, and Thimann, K. V., Am. /. Botany, 36:342 (1949)- 67. Liebscher, G., /. Landw., 35:335 ( 1887). 68. Loehwing, W. F., Proc. Iowa Acad. Sri., 49:61 (1942); Science, 107: 529 (1948). 69. Long, J. H., Proc. Am. Soc. Hort. Sci., 37:553 (1939)- 70. Luce, J. W., and Weller, R. A.. Biochem. ]., 42:412 (1948). 71. Lundegardh, H., Nahrstoff Aufnahme der Pflanze (Jena: Gustav. Fischer, 1932). 72. Mahoney, C. H., Com. Fertilizer, 65:1 (1942); Chem. Abstracts, 37: 3544 0943)- 73. Mason, T. G., Ann. Botany, 36:457 (1922); West Indian Bull., 19: 214 (1Q22). 74. ■ , and Maskell, E. J., Ann. Botany, 45:125 (1931). 75. , and Phillis, E., Ann. Botany, 48:315 ( 1934). 76. MaksTmov, N. A., The Plant in Relation to Water. Yapp, R. H., ed. and trans. (London, Allen and Unwin Company, 1929). 77. , Plant Physiology. Harvey, R. B., and Murneek, A. E., eds., Krassovsky, I. U., trans. (2d English ed., McGraw-Hill Book Com- pany. 1938). 78. McCalla, A. G., Can. J. Research, 9:542 (1933). 79. Menachem-Starkmann, T., Plant and Soil, 1 :82 (1948). 80. Mevius, W., Forschungsdienst, 16:245 (1942); Chem. Zentr., 1:2695 (1942); Chem. Abstracts, 37:4770 (1943). 81. Mitchell, J. W., Am. Naturalist, 76:269 (1942); Botan. Gaz., 99: 171, 569 (1938). 82. Mothes, K., Fortschr. Bot., 1:155 (I931)- 83. • — , Planta, 12:686 (1931). 84. , Biol. Zentr., 52:193 (1932). 85. , Flora, 128:58 (1933); Planta, 19:117 (1933). 86. , Ber. dent, botan. Ges., 51:31 (1933). 87. .Fortschr. Bot., 3:1 29 (1933); 5:202 (1936). 88. Muir, R. M., Am. J. Botany, 29:716 (1942); Chem. Abstracts, 37: ii55 (J943)- 89. Murneek, A. E., Botan. Gaz., 79:329 (1925). 90. , Plant Physiol., 1:1 (1926). 91. , Mo. Agr. Fxpt. Sta. Res. Bull, go (1926). o2. , Mo. Agr. Fxpt. Sta. Res. Bull. 106 (1927). 93. , Plant Physiol, 7:79 (1932). 94. , Science, 86:43 (1937). 356 Mineral Nutrition of Plants 95- — — Botan. Gaz., 102:269 (I94°); Proc. Am. Soc. Hon. Sri., 34: 507 (1936). 96- — — , and Wittwer, S. H., Proc. Am. Soc. Hon. Set., 40:201, 205 (T942)- 97. Neidig, R. E., and Snyder, R. S., Idaho Agr. Expt. Sta. Res. Bull. 1 (1922). 98. Neidle, E. K., Botan. Gaz., 100:607 (1939). 99. Noggle, G. R., Plant Physiol., 21:494 (1946). 100. Olson, L. C, and Bledsoe, R. P., Ga. Agr. Expt. Sta. Bull. 222:1 (1942). 101. Ovechkin, S., Compt. rend. acad. set. U.R.S.S., 26:170 (1940). 102. Overbeek, }. van, Am. J. Botany, 31:265 (1944); Am- /• Botany, 29: 677 (1944). 103. , Science, 102:621 (1945). 104. Pearsall, W. H., Ann. Botany, 37:261 (1923); Endeavour, 8:99 (1949). 105. Pember, F. R., R. I. Agr. Expt. Sta. Bull. 169 (191 7). 106. Penston, N. L., Nature, 136:268 (1935). 107. Petrie, A. H. K., Ann. Rev. Biochem., 11:595 (x942)- 108. , Biol. Rev., 18:105 (x943)- 109. Phillips, M., Goss, M. J., Davis, B. L., and Stevens, H., /. Agr. Research, 59:319 ( 1939); 64:533 (1942). no. Rankin, W. H., Soil Sci. Soc. Am. Proc, 11:384 (1946). in. Rippel, A., Biochem. Z., 187:272 (1927). 112. — , and Ludwig, O., Biochem. Z., 155:133 (1925); 177:318 (1926); 178:272 (1927); 187:272 (1927). 113. Robinson, M. E., New Phytologist, 28:117 (1929). 1 14. Sabinin, D. A., Bui. Mos\. Obsc. ispyt. Prir., 46:67 (1937). 115. Sapehin, A. A., Bull. acad. sci. U.R.S.S., Ser. Biol., 451 (1940). 116. Saskin, F. D., Trudy Konferents. Pocvoved. Fiziol. Saratov., 2:302 (1938); Compt. rend. acad. sci. U.R.S.S., 18:303 (1938); 27:1042 (1940). 117. Sayre, J. D.. Plant Physiol., 23:267 (1948). 118. Schmitt, L., and Schineis, W., Boden\unde u. Pflanzenernahr., 26:137 (T942): Chcm. Abstracts, 37:2868 (1943). 119. Schropp, W., and Arenz, B., Boden\unde u. Pflanzenernahr., 24:24 (i940- 120. Selke, W., Boden\unde u. Pflanzenernahr., 20:1 (1940). Chem. Abstracts, 37:5817 (1943). 121. Singh, B. N., and Lal, K. N., Ann. Botany, 49:219 (1935). 122. — , and Singh, R. B., /. Indian Botany Soc, 16:63 (1937); Chem. Abstracts, 34:2417 (1940). W.F.Loehwing 357 123. Sisakyan, N., and Kobyakova, A., Bio\himiya, 5:225 (1940); Biol. Abstracts, 16:8030 (1942). 124. ■ , and Kobyakova, A., Bio\himiya, 6:50 (1941); Acad. Set. U.S.S.R., Herb. Abstracts, 13:162 (1943). 125. Skaskin, F. D., Compt. rend. acad. sci. U.R.S.S., 27:1042 (1939)- 126. Skoog, F., Am. f. Botany, 27:939 (1940). 127. Smirnov, A. I., et a!., Planta, 6:687 (1928). 128. Stankov, N. Z., Doklady Vsesojuz. A\ad. Seljs\ohoz. Nau\., 13:33 (1939); Himiz. Soc. Zemled, 5:74 (i938h Herb. Abstracts, 10:148 (1940). 129. Stefanovskii, I. A., Sele{. Semenovod., p. 23 (1936)- 130. Struckmeyer, B. E., Botan. Gaz., 103:182 (1941); Biol. Abstracts, 16:5474(1942). 131. , and Roberts, R. H., Botan. Gaz., 100:600 (1939). 132. Sybil, E. W., Jr., M. S. Thesis, State Univ. of Iowa (1947)' PnD- Thesis (1949). 133. Thimann, K. V., Am. /. Botany, 25:270 (1938); Am. Naturalist, 75: 147 (1941). 134. Tsui, Cheng, Am. f. Botany, 35:173 (1948). 135. Ulrich, A., Am. ]. Botany, 28:526 (1941); 29:220 (1942); Chem. Abstracts, 36:517 (1942). 136. Vladescu, I., Bui. cultivar. jermentar. Tutunului, 287:231 (1934); Chem. Abstracts, 29:2204 (1935). 137. Wadleigh, C. H., Ar\. Agr. Expt. Sta. Bull. 446:1 (1944). 138. Wagner, H., Z. Pflanzenernahr. Diingung u. Boden\., A 25:48 (1932). Chem. Abstracts, 27:2471 (1933). 139. Ward, G. M., Dom. of Canada Dept. Agr. Publ. 729 (1942). 140. Watson, Ruth, Australian J. Exptl. Biol. Med. Sci., 17:241 (1939); 18:313 (1940). 141. Went, F. W., Am. Sci. 31:189 (1943). 142. White, H. C, Ga. Agr. Expt. Sta. Bull. 108 (1914); 114 (i9T5)- 143. Whyte, R. O., Nature, 123:413 (1929). 144. , Crop Production and Environment (London, Faher and Faber, Ltd., 1946). 1^, , and Murneek, A. E., with Others. Vernalization and Photoperiodism (Waltham, Massachusetts, Chronica Botanica Co., 1948). 146. Wiebe, H. H., M. S. Thesis, State Univ. of Iowa (1949). 147. Wilfarth, H., Romer, H., and Wimmer, G., Landw. Vers. -Sta., 63:1 (1906). 148. Williams, R. F., Petrie, A. H. K., et ah, Australian ]. Exptl. Biol. Med. Sci., 14:135, 165 (1936); i5:385 (1937); l6:65> 347 (J938); T7: 123 (i939)- 358 Mineral Nutrition of Plants 149. Wilton, O. C, Botan. Gaz., 99:854 (1938). 150. Wittwer, S. H., Mo. Agr. Expt. Sta. Res. Bull. 371 (1943); Science 98:384 (1943). 151. Wostmann, E., ]ahrb. wiss. Botan., 90:335 (1942). 152. Yapp, R. H., and Mason, U. C, Ann. Botany, 46:159 (1932). 153. Zaleski, W., Ber deut. botan. Ges., 19:331 (1901); Biochem. Z., 23: 150 (1910). 154. Andrews, W. B., The Response of Crops and Soils to Fertilizers and Manures (State College, Mississippi, 1947). CHAPTER I OCorrelations between Protein-Carbohydrate Metabolism and Mineral Deficiencies in Plants ROBERT A. STEINBERG I interpretation of the data on protein and carbohydrate metabolism of the green plants subjected to mineral deficiency has been complicated by the use of a wide range of climatic conditions, differ- ences in sampling, and the nonspecific nature of the methods of chemi- cal analysis. Practically no information exists concerning the chemical mechanisms of symptom formation with mineral deficiency. The great- est difficulty, however, can be attributed to our meager information on the organic nutrition of the plant. It is for this reason that a brief introduction dealing with protein and carbohydrate metabolism is advisable before attempting to evaluate the effects of mineral deficiency. PROTEIN AND CARBOHYDRATE METABOLISM A diagrammatic outline of the probable course of organic nutrition has been published by Gregory and Sen (14) and is here reproduced in Figure 1. Chibnall (6) has pointed out how adequately it represents the facts as far as they are known. An examination of the chemical reactions indicated in this diagram reveals that a breakdown of hexose molecules is assumed to precede the formation of organic acids. Interaction of these with ammonium ion results in production of amino acids. Protein is formed through con- densation of amino acids and these are regenerated on subsequent hydrolysis. Amides may or may not participate in protein formation. Respiration, i.e., oxidation with liberation of carbon dioxide, is repre- 360 Mineral Nutrition of Plants sented as having a multiple source, namely, breakdown of hexose, of aliphatic acids, and of amino acids. The consensus of opinion hitherto has been that the protoplasmic proteins remained unchanged after formation, that is, only the storage proteins were labile. Recent studies by Vickery, Pucher, Schoenheimer, and Rittenhouse (56) disclosed that isotopic nitrogen supplied to the fProU-'i Amide 4. v Z ,<7 inino Cud vNHy __lE= \ Amino acid ED T Carbon "^ residues Undetermined, carbohydrate _^ Oraan'tc acids'^- 3 Carbon compounds CO, I , V I Sucrose It Hexose • , 2 Carbon Compounds + CO* 1 Figure i. Diagram from Gregory and Sen (14) illustrating the probable course of protein and carbohydrate metabolism in plants. tobacco plant as ammonium salt appeared in its proteins within four hours. Calvin and Benson (5) could detect the presence of radiocarbon in the amino acids (alanine and serine) of an alga after a period of thirty seconds of photosythesis with C*Ol>. The extreme rapidity of these reactions is a striking demonstration of the manifold chemical reactions taking place in the hitherto presumably static protoplasmic proteins of the cell. This is one reason that the amino acid-protein cycle depicted in the above scheme seems to the writer to be the weakest part of the picture. Alternate condensation and liberation of amino acids would seem a rather futile procedure; especially so since there is some evidence that protein may participate in the formation of respiratory carbon dioxide. The writer is of the opinion that many of the metabolites of the cell, including other proteins and enzymes, are formed and liberated through cleavage from the permanent or protoplasmic proteins. Only the debris of these reactions would undergo complete oxidation to Robert A. Steinberg 361 carbon dioxide and ammonia, or be reworked to carbohydrate or other compounds. This interpretation would be in better agreement with the surmise of Gregory and Sen that breakdown of protein follows a different course than synthesis. Another objection to a simple regenerative amino acid-protein cycle in plants is based on the phenomenon of "nitrogen balance" in animals. A daily loss of amino nitrogen must be compensated in the diet, since it persists even during extreme starvation. The "wear and tear" protein metabolism is accompanied by loss of amino nitrogen as urea or uric acid, even in the starving animal. It seems logical to the writer, there- fore, to assume that a similar cycle of nitrogen and carbon loss occurs in the plant. While traces of urea and uric acid have been identified in plants, respiratory carbon dioxide probably accounts for the carbon while the nitrogen is reutilized to form new amino acids. Vickery and others (56) have remarked on the close similarity between "nitrogen balance" in animals and the rapid intake by tobacco proteins of radio- nitrogen, though Vickery (55) categorically later denied the existence of "nitrogen balance" in plants. It is true that technically "nitrogen balance" as such does not exist in plants since nitrogen is retained and re-used. However, in the larger phenomenon of which it is only one facet, it must be assumed that loss of carbon and nitrogen from proto- plasmic protein is inherent, automatic, and irreversible in all cells. It is truly cyclic. Other data bearing on the cyclic interpretation of protein metabolism consist in the fluctuations of protein and carbohydrate in plants. Protein varies little in comparison to carbohydrate, if storage products of both are omitted from consideration. Brief starvation experiments readily demonstrate this difference. Protoplasmic protein can be depleted or altered to such an extent as to prevent resumption of growth only through starvation of long duration. This is not the case, within limits, with carbohydrate. Increases in carbohydrate above minimal have little influence on protein formation, whereas increases in nitrogen (particu- larly as ammonium salts) do and may lead to marked depletion of car- bohydrate. It might also be mentioned that carbohydrate supplied externally to the green plant is utilized efficiently, whereas amino acids so supplied are usually harmful even in high dilution (49). Finally 362 Mineral Nutrition of Plants the so-called "sparing action" of sugar on respiratory destruction of amino acids (27) is in all probability a "replenishment action" instead. These and other data and interpretations will be found in reviews by Nightingale (29, 30), Chibnall (6), and Vickery (54). The writer has depended to a great extent on these and other workers for information of the accomplishments in the field of protein chemistry. A brief reference might be made to the question of amino acid synthesis concerning which so little is known. It was possible by means of Aspergillus niger van Tiegh, a fungus, to demonstrate that normal and optimum growth could take place only with a mixture of glutamic acid, proline, and ornithine in the absence of sugar and other nitrogen (46, 4J). The other amino acids, excepting aspartic, were ineffective, whereas malic acid was also usable. The aspartic and malic acids have the same carbon chain. Adoption of a classification of amino acids into three groups has received some measure of verification through the work of Steward and Street (52) and of MacVicar and Burris (26) on the importance of glutamic acid in amino acid synthesis. Stepka, Benson, and Calvin (5/) found serine and alanine formed during photosynthesis to be radioactive, whereas the large quantities of gluta- mic acid elaborated were inactive. Synthesis of straight chain amino acids with 3 and 4 carbon atoms would seem to follow a different path than those with a 5 carbon atom chain according to their data. THE DATA OF MINERAL DEFICIENCY It is evident in view of the speed and complexity of nitrogen and carbon metabolism in plants that any deficiency of a mineral element participating directly in protein and carbohydrate reactions would be readily indicated through chemical analysis. The same would also be true for an indirect participation, as in enzymes controlling these reac- tions. However, deficiencies drastic enough to influence a basic reaction of protoplasm would affect all physiological processes and could not be accepted as proof of a role in a limited phase of nutrition. A feature of mineral nutrition studies that has long puzzled the writer has been the chemical mechanism whereby characteristic defi- ciency symptoms are produced in the plant. Data obtained in aseptic culture with amino acids has indicated the probable participation of Robert A. Steinberg 363 these compounds in these phenomena (48, 49). These data are included at the end of this paper. Certain assumptions and simplifications have been introduced in the data of the literature here reproduced. All values have been recomputed in terms of "deficiency/control," i.e., content of a constituent in the deficient plants divided by that in the controls. Variations as between chemical methods so tend to be minimized. Carbon values are limited to total carbohydrate exclusive of hemicellulose and cellulose, and its major cleavage product— reducing sugar. Nitrogen values are similarly limited to protein and the major cleavage product— amino acid. Values for soluble nitrogen have also been used, particularly if corrected for nitrate. Four points have been kept in mind by the writer in the evaluation of the numerical data, (a) The selected data and the conclusions they entail concerning the individual researches are those of the investigators whose papers are cited, (b) The data in these publications have been recomputed for use in these tables and should be compared on a relative and individual basis. That is they should be compared in pairs: free amino nitrogen versus protein, and reducing sugar versus carbohydrate. While many of the tabulated values are also true on an absolute basis, this is considered relatively immaterial from the viewpoint of consecu- tive chemical reactions, (c) Mild or no symptoms are given equal weights in the means and the total number of positive results despite any proof other than the intention of the investigator that a deficiency existed, (d) The calculated means and the number of positive results are used as a rough measure of the extent of agreement in investiga- tions. However, the degree of mineral deficiency, the magnitude of differences in pairs of values, and the extent and consistency of the values in each investigation are also used as aids in interpretation. EFFECTS OF MINERAL DEFICIENCIES ON THE PROTEIN AND AMINO ACID CONTENT OF PLANTS A summary of the results reported by investigators as concerns the effects of mineral deficiency on protein metabolism is contained in Tables I— III. These results are limited on the whole to leaves and stem. Special procedures of sampling such as the use of fractional parts of the 364 Mineral Nutrition of Plants stem are so indicated and enclosed in parentheses; analyses of roots or fruit are also placed in parentheses. Table I deals with the effects of nitrogen, phosphorus, and potassium deficiency. The values for amino nitrogen in the leaves of nitrogen- deficient plants seem to indicate a slight tendency to increase as com- pared with the controls. If these ratios are averaged, however, it is seen that the amino acid content of deficient plants is identical with that of normal plants, while protein has decreased by one-third. On a relative basis, however, ami no-nitrogen has increased as compared with protein. Analogous values for the stems differ only slightly. Richards and Templeman (j6) interpret their data on nitrogen deficiency to mean that no marked disturbance of protein metabolism takes place. The quantitative data tabulated for nitrogen deficiency verify the earlier conclusions of Kraybill and Smith (2^) and Kraybill (22). An interest- ing study of nitrogen deficiency in corn grown with ammonium-nitro- gen (57) should be consulted. Wood and Petrie (6/) conclude from their data that the concentration of sugar applied from an external source did not disturb the relation of amino acids to protein in leaves of Phalaris tuberosa L. Sucrose, glucose, and fructose were used. Results on phosphorus deficiency are less ambiguous. A definite in- crease in amino acid nitrogen and decrease in protein is readily evident both in the leaves and stems. This interpretation is supported by the values of the means and again agrees with the interpretation of Richards and Templeman (j6). The reasons for the apparent variations in the results of different investigators are not entirely clear, except that those data obtained with material from plants showing severe symptoms of phosphorus deficiency show the greater differences. The phosphorus deficiency data are also in accord with the prior work of Kraybill and Smith (2j) and Kraybill (22) on tomato plants. Those of Turtschin (55) and of Smirnov et al. {44) also show good agreement. Potassium deficiency studies seem to have been favored by investiga- tors. These have reported, with few exceptions, that insufficiency of potassium causes a disturbance in protein metabolism indicated by a relative increase in amino acid nitrogen and a decrease in protein. Both leaves and stems are affected. Richards and Templeman (j6) interpret these results as due to breakdown of protein and not an inhibition of TABLE I Effects of Mineral Deficiency on Free Alpha-Amino Nitrogen and Protein Nitrogen of Plants: Nitrogen, Phosphorus, Potassium Symp- Deficiency/C ontrol Ratios Plant Leaves Stems Refer- toms Amino Protein Amino Protein ence acid-N* N acid-N* N Nitrogen deficiency Tung mild 0.67 0.77 (8) Barley medium 1.38 less Oi) Tomato 0.21 0.25 0.18 0.25 (ij>) Barley severe 1. 41 0.95 (36) Tobacco severe 1.32 (50) Mean 1. 00 (5) 0.66 (j) 0.18 (/) 0.25 (/) Phosphorus deficiency Tomato (NO3 N) mild 0.90 1.05 0.95 1.36 (2) Tomato (Urea -N) mild 0.74 0.86 1. 13 0.80 (2) Barley medium 1 . 10 less {'3) Tomato 0.99 0.78 1.80 0.40 ('9) Wheat (tops) severe (i.o9)s (o-94) (2J) Wheat (roots) severe (1.48)° (0.78) (2/) Wheat (seed) severe (1.28)3 (0.97) («) Toma to medium 0.533 0.83 0.838 1. 14 (^5) Barley severe 2.09 0.66 (36) Tobacco mild 1.48 (5c) Mean .12(7) 0.84(5) 1. 18(4) 0.93(4) Potassium deficiency Soybean medium 1.3! 0.90 1. 14 0.72 M Guayule (immature) medium 2 40 1.24 (7) Guayule (mature) medium 2 98 I23 (7) Tung mild 0 83 1 .01 (*) Barley medium 0 64 more (•3) Sugar cane medium 3 3° 0.92 1 . 12 0.92 ('6) Covvpea mild 1 09 0.94 0.95 1. 18 (,8) Sugar beet mild 1 43 i-39 (,S) Sugar beet (roots) mild (1 12) (1.10) (18) Tomato 2 758 1 .04 2.32s 0.85 ('9) Tomato medium 1 38 1 .02 1 .32 1.07 (3^) Tomato mild 1 29s 1 . 1 1 1 .23s 1 .00 (35) Barley medium 1 3° 0.92 (36) Oats medium 2 913 1.46s (37) Pineapple (NO3 -N) 1 23 1 .09 1 .22 0.80 (42) Pineapple (NIT -N) 2 46 1. 31 1.82 0.73 (42) Tobacco severe 6 87 (50) Tomato medium 1 92s0 1 .02 1.73s0 I .21 (59) Mean 2'3 {'?) 1 .04 (/6) 1 43 (9) O.94 (9) ''Values representing soluble nitrogen are indicated by "s," and soluble organic nitrogen by so. 366 Mineral Nutrition of Plants synthesis. Other data (50) clearly indicated that the magnitude in the difference obtained depends on the severity in deficiency symptoms and the degree of exclusion of apparently normal tissues. That is to say, that both the symptoms and the degree of chemical change are local- TABLE II Effects o Mineral Deficiency on Free Alpha-Amino Nitrogen and Protein Nitrogen of Plants: Magnesium, Calcium, Sulfur Symp- toms Deficiency/Control Ratios Plant Leaves Stems Refer- ence Amino acid-N Protein N Amino acid-N Protein N Magnesium deficiency Soybean Tobacco severe 1 .36 3.83 0.85 0.71 0.91 to (50) Mean .60(2) 0.85(7) 0.71(7) 0.91(7) Calcium deficiency Soybean mild 0.63 0.65 0.68 0.67 to Tomato (-Cu also) 0.93 1.03 (3') Tobacco severe 2.20 (50) Mean 1.42 (2) 0.65 (7) 0.81 (2) 0.85(7 Sulfur deficiency Soybean medium 1. 15 0.87 2.61 1. 51 (9) Soybean (roots) medium (2.02) (0.71) (9) Sunflower (upper) medium 4-31 0.85 (ro) Sunflower (middle) medium 5-43 0.71 (jo) Sunflower (lower) medium 2-74 0.77 (10) Tomato slight 3.88 0.90 (33) Mean 1.15(7) 0.87(7) 3-79 (5) 0-99 (5) ized. Nevertheless, Phillips, Smith, and Dearborn (_#) were unable to find any indication of derangement of nitrogen metabolism with potas- sium-deficient tomato plants. Deficiences in magnesium, calcium, and sulfur (Table II) seem to have an identical influence on the relative quantities of amino acid nitrogen in plants. Most investigations agree in reporting an increase in free amino acids in the leaves of plants deficient in either magnesium, calcium, or sulfur. However, no significant change is reported by Robert A. Steinberg 367 Burrell (4) as taking place in the stems of soybean with either calcium or magnesium deficiency. Nightingale, Addoms, Robbins, and Scherm- erhorn (•?/) also found little alteration in amino acid and protein con- tent of tomato stems to occur with calcium deficiency. The data for sulfur deficiency in stems showed more differences in all cases in regard to both chemical reactions. The reason for the unchanged values in stems with magnesium and calcium deficiency might be one of several. It may be a characteristic reaction to a deprivation of these elements. A mild deficiency could also possibly cause a similar condition. Further, it should be recognized that, if localization of symptoms is paralleled by alterations in amino acid and protein content, the stems of plants should be relatively less affected than the leaves. Skok (43) obtained calcium deficiency symptoms with the bean plant when supplied with urea as a source of nitrogen. Additional studies would seem desirable, but it seems clear that nitrate reduction is not a primary function of calcium, nor the inactivation of nitrate a causative factor in calcium deficiency. The data on the effects of deficiencies in micronutrients are very few (Table III). Bennett (/) reports that iron deficiency leads to an in- crease in amino acid nitrogen and a decrease in protein nitrogen. Gilbert, Sell, and Drosdoff (12) reported a definite increase in both constituents in the leaves of tung during early stages of growth with copper defi- ciency. Later stages showed a marked decrease in free amino acids, however. These results may afford an explanation of those reported by Lucas (24), who concluded that copper does not participate in protein metabolism. The increased protein may well be due to an inhibition in protein breakdown, a view in accord with that of Gregory and Sen (14) that synthesis and breakdown of protein followed different paths. Cop- per enzymes may play a relatively small part in the former reaction as compared with the latter. The data of Mulder (28) on molybdenum deficiency are different for leaves and stems. This author emphasizes the decrease in protein ac- companying a molybdenum deficiency and the large increases in un- utilizable nitrate. The leaves show a relatively lesser decrease in amino acids than in protein. Insufficient data on the effects of boron deficiency also do not permit 368 Mineral Nutrition of Plants TABLE III Effects of Mineral Deficiency on Free Alpha-Amino Nitrogen and Protein Nitrogen oi Plants: Iron, Copper, Molybdenum, Boron Plant Symp- toms Deficiency/Control Ratios Leaves Amino acid-N* Protein N Stems Amino acid-N* Protein N Refer- ence Iron deficiency -f — Copper deficiency Tung (early) mild i 89 1.46 Tung (late) mild 0 43 1.32 Alfalfa 1 .09 Barley i-43 Oats 1. 13 Wheat 1 .04 Wheat (grain) (1. 11) Sugar beet 1 .23 Tomato (fruit) (114) Carrot (root) (1.48) Mean 1.16(2) 1.24(7) (') (12) (™) {24) (24) (24) (24) (24) (24) (24) (24) Molybdenum deficiency Tomato mild 0.96 0.83 Tomato (roots) (0.84) (1.07) Mean 0.61 0.78 0.96(7) 0.83(7) 0.61(7) 0.78(7) (28) (28) oron deficiency Nasturtium medium 0 . 70s0 0 ■72 Nasturtium (roots) (0.65)80 (0 .76) Alfalfa (tops) medium (1.69)80 Spinach mild (i.o6)s (0 89) Tobacco mild 1.27 Mean 1.088 0.78 0.99(2) 0.72(7) 1.08 (/) 0.78(7) (J) (3) (38) (39) (5o) ''Values representing soluble nitrogen are indicated by "s," and soluble organic nitrogen by a clear cut interpretation. Scripture and McHargue reported increased free amino acids in alfalfa tops (38) and spinach (jo) and a drop in protein in the latter. Briggs (3) found amino acid and protein nitrogen decreased to about the same extent in the leaves of the nasturtium plant, whereas the former increased and the latter decreased in the stems. A Robert A. Steinberg TABLE IV 369 Effects of Mineral Deficiency on Reducing Sugars and Total Carbohydrates (sugars plus starch) of Plants: Nitrogen, Phosphorus, Potassium Symp- toms Deficiency/Control Ratios Plant Leaves Stems Refer- Reducing sugar Carbo- hydrate* Reducing sugar Carbo- hydrate* Nitrogen deficiency Tung Barley Tomato mild severe 0.89 1.08 0.36 0 93 1.78s '■55 0.59 0.93 ('J) (>9) Mean 0.78 (3) 1.24(2) 0.59(7) 0.93(7) losphorus deficiency Barley mild i-33 l.IO8 ('?) Lemon °-73 0.51" 0.14 0. 19 ('5) Tomato medium 0.24 1 .09 0.87 0.95 (19) Wheat (dark-7 days) 1 .29 1.29 («) Wheat (root-seed) (o.79) (0 ■ 79) (1.35) (1.01) (27) Tomato (average) medium 1.63 0.86 1.79 i-34 (25) Mean .04(5) 1.08 (3) 0.93 (3) 0.69 (3) Potassium deficiency Soybean 0.83 1.24 0.58 0.64 w Guayule (immature) medium 2.39 i-47 3.46 1 .09 (7) Guayule (mature) medium 3-41 1 .06 0.96 0.63 (7) Tung moderate : 0.75 0.90 (*) Barley mild 0.80 0.638 ('J) Sugar cane (9 week) medium 1 . 10 0.968 0.85 1.08s (76) Sugar cane (7^ month) medium 1. 17 1 . 028 0.43 0.65s (76) Pea mild 0.67 1.03 0.80 0.90 ('?) Cowpea 1.28 1 . 1 1 1.62 1.27 (/*) Sugar beet (roots) 0.80 0.87 (0 . 70) (0.65) (/*) Tomato mild 0.66 0.62 0.86 0.83 ('9) Tomato medium 0.58 0.47 1. 16 0.75 {32) Tomato 0.91 1 . 10 1.24 i'3 (55) Pineapple (NO3-N) 1 .96 1. 41 1. 17 0.66 U') Pineapple (root) (3- ") (1.84) (4') Pineapple (NH4-N) 2-35 0.92 i-47 0.61 (4') Pineapple (root) (i-i9) (0) (4') Tomato severe 1 .76 i-33 1. 31 0.81 (59) Tomato (nitrate) medium 1 .09 1 .09 (60) Tomato (ammonia) medium 0.84 0.82 (60) Mean 1.34 (/6) 1.03 (75) 1.22 {13) 0.86 (13) *Values for total sugar are indicated by "s." 37° Mineral Nutrition of Plants parallelism between severity in symptoms and degree of chemical change has been noted for boron by Steinberg, Bowling, and McMurtrey (50). A microchemical study by Wadleigh and Shive (5$) led these investiga- tors to conclude that boron deficiency caused an alteration in the normal course of protein synthesis in cotton seedlings. EFFECTS OF MINERAL DEFICIENCIES ON THE REDUCING SUGAR AND CARBOHYDRATE CONTENT OF PLANTS Nitrogen, phosphorus, and potassium deficiencies in Table IV also reflect the effects of divergencies in sampling and climatic conditions. However, certain variations in composition have been noted by in- vestigators in their experiments. These changes consist, with nitrogen deficiency, of a relative decrease in reducing sugar and an increase in total carbohydrate. The increase in carbohydrate is attributed to dimi- nished formation of protein. Phosphorus deficiency leads to equal in- creases in both fractions for the same reason, while lack of potassium causes a greater increase in reducing sugar than in total carbohydrate. The mean values for nitrogen deficiency in Table IV indicate a relative decrease in reducing sugar as compared with total carbohydrate both in leaves and stems. Reducing sugar apparently increased rela- tively to carbohydrate only in the stems of plants subject to phosphorus deficiency, but not in the leaves. MacGillivray's data (Table VII) would indicate this to be only an apparent exception. With potassium defi- ciency, reducing sugars again seemed to show a relative increase as compared with carbohydrate in both stems and leaves on the basis of the mean values. Analytical data on the contents of reducing sugar and carbohydrate are very limited in the cases of magnesium, calcium, and sulfur defi- ciencies. Table V presents the available data. Almost invariably there is a decrease in both fractions in both leaves and stems with these three elements. Only in the case of stems short of calcium is there a slight increase in both reducing sugar and total carbohydrate. Micronutrient deficiency effects on reducing sugar and carbohydrate have been tabulated in Table VI. Iron-deficient pineapple plants are lower in total carbohydrates than are the controls and this is accom- panied by a slight relative increase in reducing sugar. Both leaves and / Robert A. Steinberg TABLE V 371 Effects of Mineral Deficiency on Reducing Sugars and Total Carbohydrates (sugars plus starch) of Plants: Magnesium, Calcium, Sulfur Symp- Deficiency/C ontrol Ratios Plant Leaves Stems Refer- ence Reducing sugar Carbo- hydrate Reducing sugar Carbo- hydrate Magnesium deficiency Soybean mild 0.88 0.85 0.73 0.74 U) Mean 0.88 0.85 0.73 0.74 Calcium deficiency Soybean Pea Tomato 1 .11 0.47 1.08 0.70 1 . 10 0.40 1.94 0.91 0.72 2.07 (4) ('?) (3') Mean 0.79(2) 0.89(2) 1.15(3) !-23(i) Sulfur deficiency Soybean Soybean (roots) Sunflower Mean medium 0.84 0.84 medium (0.67) (1.20) 0.84 0.84 0.18 1.44 0.23 0.27 0.21 (2) 0.86 (2) (9) (9) (10) stems responded similarly. Copper deficiency had a similar effect on leaves of the tung tree. Withholding boron, however, led to increases in total carbohydrates in leaves and stems, with relatively increased reduc- ing sugar in the former and a decreased content in the latter. SOME DATA OF INDIVIDUAL INVESTIGATIONS A clearer idea of the meaning of the preceding summaries can be ob- tained through an examination of some of the specific data. The follow- ing have been selected on the basis of consistency in results, size of plant samples, and comparison of effects in different organs in the plant. Data meeting these criteria are confined almost entirely to nitrogen, phosphorus, and potash. Attention is again called to the fact that these data have been recomputed in terms of "quantity in deficient plant/quantity in the control." 372 Mineral Nutrition of Plants TABLE VI Effects of Mineral Deficiency on Reducing Sugars and Total Carbohydrates (sugars plus starch) of Plants: Iron, Copper, Boron Plant Symp- toms Deficiency/C ontrol Ratios Leaves Stems Reducing sugar Carbo- dyhrate* Reducing sugar Carbo- hydrate Refer- ence Iron deficiency Pineapple (NO3-N) 1 .04 0.89 0.90 0.84 (40) Pineapple (NH4-N) 0.74 0.80 1 .02 0.77 (40) Mean 0.89 0.85 0.96 Copper deficiency Tung (early) mild 0.77 0.79 Tung (late) mild 0.97 0.83 Mean 0.87 0.81 (/2) Boron deficiency Tomato medium 2.14 i-73 Tomato (roots) medium (2.09) (1.85) Alfalfa (tops) medium (1 .85) (1.66) Mean 14 i-73 0.61 0.61 1 .06 1 .06 {20) {20) *Values for total sugar indicated as "s." The effects of phosphorus deficiency on carbohydrate metabolism of the tomato plant obtained by MacGillivray (25) are illustrated in Table VII. Phosphorus deficiency caused a relative increase in reduc- ing sugars in both leaves and stems. Whereas total carbohydrates — sugars plus starch — increased to a somewhat lesser extent also in the stems, a definite decrease took place in the leaves. There was on the whole a tendency for all carbohydrate to concentrate in the lower portions of the plant. Some degree of correlation exists therefore be- tween the loci of symptoms and of greatest variations in reducing sugars and total carbohydrates. The data of Nightingale, Schermerhorn, and Robbins (32) on the effects of potassium deficiency on the tomato plant are shown in Table VIII and include values for amino and protein nitrogen. Reducing Robert A. Steinberg 373 TABLE VII Effect of Phosphorus Deficiency on the Carbohydrate Content of the Tomato Plant* Plant Part Deficiency/Control Ratios: Phosphorus Reducing sugar a rbohydratesf 1.16 1.03 1.52 2.82 0.71 0.45 0.74 1.54 Leaves Top Next to top Middle Next to bottom Mean 1.63 0.86 Stems Next to top Middle Next to bottom Bottom 1.65 1.86 1.99 1.67 0.98 1.28 i-55 i-55 Mean ■79 !-34 *Computed from data of MacGillivray (25) jSugars plus starch TABLE VIII Effect of Potassium Deficiency on Protein and Carbohydrate Content of the Tomato Plant* Defici ency/Control Ratios: Potassium Reducing Carbo- Plant Part Amino-N Protein-N sugar hydrate Upper blades '•34 0.97 0.39 0.39 Lower Blades 1. 41 1 .07 0.60 0.48 Upper petioles 2-57 0.92 0.67 0.67 Lower petioles 1 .24 0.71 0.64 0.34 Upper stems '•35 0.98 1 .07 0.83 Lower stems 1 .29 1. 16 1.25 0.66 Roots 1 .60 1 .00 0.97 0.84 ^Computed from data of Nightingale, Schermerhorn, and Robbins (32) 374 Mineral Nutrition of Plants sugars in the deficient plants decreased in the leaves and increased in the stems. Total carbohydrate gave consistent decreases in all parts of the plant. The net result was a relative increase in reducing sugars as compared with total carbohydrate because of potassium deficiency. Carbohydrates tended to increase in the lower parts of the plant, where symptoms of potassium deficiency are most extreme. Protein content also seems to become higher in the lower leaves and stems of plants lacking in potassium. Amino nitrogen shows a definite increase as compared with normal plants, due either to proteolysis or more probably to a block in protein formation. The quantities of all constituents parallel each other in kind as between upper and lower leaves. The effects of deficiency in causing a relative increase in break- down products — amino acids and reducing sugars — are plainly visible in the roots. The sampling procedure followed by Richards and Templeman (^6) and Gregory and Baptiste (/j) depended on the analysis of successive leaves of barley plants as each reached maturity. The effects of defi- ciencies would be expected to increase in severity with successive leaves unless the element lacking was rapidly and completely mobile. Examination of the data of Richards and Templeman (j6) in Table IX discloses that the relative quantities of amino and protein nitrogen remained unaltered by nitrogen deficiency; whereas the former showed marked increases with phosphorus and potassium de- ficiencies. These displacements in relative quantities were caused in part by a relative fall in protein with insufficient phosphorus. Protein ap- parently did not decrease with potassium deficiency. These authors concluded that phosphorus was necessary for the formation of protein, and that potassium did not participate in this process. The increased amino acid content of leaves lacking potassium in barley was attributed to proteolysis. The results for carbohydrates with barley plants subject to nitrogen, phosphorus, and potassium deficiency have been reported by Gregory and Baptiste (/j). Some of their data have been tabulated in Table X. Total carbohydrates were not determined, unfortunately. Total sugars seemed to show an increase with deficiencies in nitrogen and phos- phorus, but to diminish sharply with lack of potassium. TABLE IX Effect of Nitrogen, Phosphorus, and Potassium Deficiency on the Protein Content of Consecutively Matured Leaves of Barley* Deficiency/Control Ratios Leaf Number from Base to Tip Low nitrogen Low phosphorus Low potassium Amino- N Protein- N Amino- N Protein- N Amino- N Protein- N i I .00 O.99 1 . 11 0.97 1 .00 0.97 2 0-99 I .00 0.98 0.78 1 .01 0.84 3 O.96 1 .40 1 .04 1 .42 1 . 11 1. 51 4 O.90 0.95 0.90 0.91 0.97 0.95 5 0.68 0.50 I-55 0.72 1 .22 0.87 6 °-53 0.41 1.32 0.64 1.30 0.96 7 0.40 0.29 2. 16 0.69 1 55 0.86 8 0.47 0.38 1.58 °-93 1.58 1 . 10 9 0.52 0.53 1 54 0.85 2.04 0.95 10 — 0.57 3-24 0.94 3.01 1 . 10 Mean 0.72 0.70 1 54 0.89 1.48 1 .01 *Computed from data of Richards ; uid Templeman (j6) TABLE X Effect of Nitrogen, Phosphorus, and Potassium Deficiency on the Carbo- hydrate Content of Consecutively Matured Leaves of Barley* Leaf Number from Base to Tip Deficiency/Control Ratios Low nitrogen Reduc- ing sugar Sugar Low phosphorus Reduc- ing sugar Sugar Low potassium Reduc- ing sugar Sugar I 1 .20 0.92 1.08 0.90 1.08 0.95 2 1. 18 2.15 0.97 i-93 0.82 0.94 3 0.99 1 .06 0.94 1 .01 1. 14 0.87 4 1 .02 i-23 0.86 0.72 0.85 0.43 5 1 . 10 2.01 i-45 1. 41 0.98 0.89 6 0.67 1 .76 1.27 0.57 0.49 0.21 7 — — — — 0.54 0.38 8 0.52 1. 16 0.59 0.66 °-35 0.30 9 1. 14 1.98 1. 16 o-93 0.79 0.59 10 1.59 1 .04 ^^~ ~~ — Mean 1.05 1.48 1 .04 1 .02 0.78 0.62 ^Computed from data of Gregory and Baptiste (/j) 376 Mineral Nutrition of Plants The data obtained by Wall (59) on potassium shortage of the tomato plant (Table XI) agree with those of Richards and Templeman in Table IX in regard to the disproportionate increase in amino acids and maintenance in protein content. They differ from those of Gregory and Baptiste (Table X) in showing both an increase in total carbohydrate and the relatively increased content of reducing sugars. Wall pointed TABLE XI Effect of Potassium Deficiency on the Protein and Carbohydrate Content of the Tomato Plant* Plant Part Deficiency/Control Ratio: Potassium Soluble Organic-N Protein-N Reducing Carbo- hydratef Leaves Upper Lower Mean Stems Upper Lower Mean 1.86 1.98 1 .92 1 .90 1.56 r-73 1. 16 0.87 1 .02 1 -37 1 .04 1 .21 1.03 3-93 2.48 0.69 1. 16 °-93 0.87 2.85 1.86 0.72 0.55 0.64 *Computed from data ot Wall (59) fSugars plus starch out, therefore, that the evidence favored potassium participation in protein formation, since proteolysis does not ordinarily occur in the presence of adequate sugar. In these experiments sugars and carbohydrates were higher in the lower parts of the plant, while amino acids and protein tended to be greater in the younger tissues. The nitrogen differences were slight, however. The experimental results of Sideris and Young have interested the writer for they afford a comparison between the effects of ammonium and nitrate nitrogen on potassium deficiency. The analytical results Robert A. Steinberg 377 for potassium deficiency in the pineapple are shown in Table XII (4/, 42). Here again will be seen a relative increase in amino acids and reducing sugars in all parts of the plants supplied with nitrate nitrogen or with ammonium nitrogen. Protein has increased in the leaves but diminished in the stems. It may be noted here that the assumption TABLE XII Effect of Potassium Deficiency on the Protein and Carbohydrate Content of the Pineapple* Deficienc\ '/Control Ratios: 'otassiurr i Plant Part Nitrate-N Ammonium-N Amino- N Protein- N Reduc- ing sugar Carbo- hv- dratef Amino- N Piotein- N Reduc- ing sugar Carbo- hy- dratef Leaves Young 1. 13 1 .09 1.63 1. 17 2.13 1 . 12 1.03 0.89 Active i-43 1. 19 J-93 i-74 2.20 1. 18 1.05 1 . 12 Mature 1. 41 1 . 1 1 2.59 1.56 2-93 1. SO 1.39 1 . 10 Old 1.23 0.94 1.92 1. 51 2.52 i-32 1.30 1 . 12 Mean 1.30 1 .09 2.02 1.50 2.46 I-3* 1. 19 1 .06 Stems Apex Middle Base 1.38 1 .22 1 .09 0.79 0.74 0.87 1.88 2.31 2.06 2.20 0.56 o-37 2.05 1.98 1.39 0.57 0.74 0.86 1 . 10 1. 81 i-74 2.13 0.61 0.60 Mean 1.24 0.80 2.08 1 .04 1.82 0.73 1 .46 1 .11 Roots 1 -73 0.80 3" 1.78 1 .02 0.78 1. 19 — *Computed from data of Sideris and Young (41, 42) fSugars plus starch that potassium deficiency owes its action to a "locking" of nitrate (//) is not borne out by the appearance of potassium deficiency symptoms in the absence of nitrate nitrogen. Definite gradients in ratios may be noted in the pineapple plant that has a potassium deficiency. The values for young and old leaves tend to show a lesser increase in nitrogen and carbohydrate constituents. Amino acids and carbohydrates decrease from the apex to the base of the stem. 378 Mineral Nutrition of Plants The data in Table XIII are included to illustrate the relative increase in reducing sugar as compared with total carbohydrate in the iron deficiency of the pineapple plant. A similar response was obtained with ammonium and nitrate nitrogen. TABLE XIII Effect of Iron Deficiency on the Carbohydrate Content of the Pineapple5 Leaves Young Active Mature Old Mean Deficiency/Control Ratios: Iron Plant Part Nitrate-N Ammonium-N • Reducing sugar Carbo- hydrate! Reducing sugar Carbo- hydrate! 1 .06 0.93 0.78 0.87 0.92 0.80 0.63 0.67 0.85 0.79 0.84 0.82 1 .29 1 .03 1 .06 0.84 1.03 0.89 ■83 0.80 Stems Apex 0.86 0.66 1 . 10 0.82 Middle 0.92 0.89 1 .04 0.69 Base o-93 °-93 0.94 0.79 Mean 0.90 0.83 1.03 0.77 Roots 0.67 1 .02 0.76 0.75 *Computed from the data of Sideris and Young (40) jSugars plus starch CHEMICAL MECHANISM OF VISUAL SYMPTOM PRODUCTION The writer has reported (48) that characteristic responses in gross morphology of tobacco seedlings take place in aseptic culture on a complete mineral agar with traces of each of the free amino acids. Isoleucine in quantities of 20-100 p.p.m. caused production of french- ing-like symptoms including network chlorosis, inhibition of stem and branch elongation, abnormal increase in leaf number, and inhibition of Robert A. Steinberg 379 leaf lamina growth. Figures 2 and 3 illustrate frenching of tobacco in the field and the isoleucine effect in aseptic culture, respectively. Hydroxyproline at 5 p.p.m. quickly killed. Various types of chloroses, necroses, and abnormalities in leaf form were characteristic of other Figure 2. Field of tobacco in Lancaster County, Pennsylvania, showing extreme frenching. — Courtesy of Dr. O. E. Street. amino acids. The conclusion was drawn that production of symptoms in the plant was probably due to abnormal protein metabolism and the excessive accumulation of metabolites. A later paper (49) confirmed and extended these results to possible aliphatic acid metabolites. Only the natural isoleucine was found to be effective, and not the unnatural optical isomer (Figure 4). A still later paper (50) dealing with field plants of tobacco showing symptoms of frenching and mineral deficiencies affords additional evidence for this interpretation. Symptoms of frenching and of calcium, magnesium, potassium, phosphorus, and boron deficiency were ac- companied in each case by marked increases in free amino acid content of the leaf lamina. These increases ranged up to about 600 per cent: 38o Mineral Nutrition of Plants Figure 3. Maryland Medium Broadleaf tobacco in asceptic culture il- lustrating the effect of 200 p. p.m. of synthetic isoleucine. nitrogen, 32; phosphorus, 48; potassium, 587; calcium, 120; magnesium, 283; boron, 27; and frenching, 94. Chloroses were attributed primarily to the accumulation of toxic metabolic products and were not, except possibly magnesium, evidence of a function in chlorophyll formation. An important corollary of these results has a direct bearing on the Robert A. Steinberg 381 "":':":?-':':'0 J. :J.S;.;: !■;■:.■ Figure 4. Maryland Medium Broadleaf tobacco in asceptic culture il- lustrating the effect on growth of 100 p. p.m. of natural isoleucine. The "D( — )" designation is an error. It should be "L(-|-)." preceding data concerned with the metabolic changes accompanying mineral deficiencies. Alterations in metabolism of the leaf lamina were found to be proportional to the degree of abnormality. Localization, it was surmised, operates even in individual leaves, the healthy portions showing relatively little change compared to abnormal portions. 382 Mineral Nutrition of Plants DISCUSSION The summary of data in the literature concerning the effect of mineral deficiency on amino acids and protein shows how greatly the results of various investigators have differed. A certain degree of uni- formity has been introduced, perhaps unwarranted, by comparing the mean values obtained in all experiments. Another procedure, namely, comparison of the number of positive results with negative results, might also have been used. A brief computation will disclose that six out of nine of the investigators with phosphorus, and fifteen of seven- teen with potassium obtained an increase in amino acids and a relative decrease in protein of the leaves with deficiency. Sulfur deficiency re- sulted in similar reactions without exception. Since the greater num- ber of experiments deal with potassium and phosphorus in the order given and only limited information is available for the other elements, these may be taken to indicate that any mineral deficiency results in a disturbance of protein metabolism. Only in a few instances did the analytical data indicate an unchanged proportion of amino acids to protein in either leaves or stems as a result of deficiency. This interpretation is also supported by the data of those experimental results of greater uniformity due in part to the use of large samples — for example, those of Sideris and Young {40, 4/), Nightingale et al. (j/, ^2) and Wall (59, 60). Steinberg, Bowling, and McMurtrey ($0) were able to demonstrate that marked increases in free amino acids took place in leaves of field-grown tobacco. Large samples of 50 to 100 leaves were used and positive results were obtained in each case (calcium, magnesium, potassium, phosphorus, nitrogen, and boron) in accordance with the severity of symptoms displayed. The mean values for nitrogen, phosphorus, and potassium deficiency in successive leaves as they matured were determined by Richards and Templeman (j6) and are shown in Table X. Nitrogen deficiency did not alter the relative proportion of amino acid and protein. Phosphorus deficiency caused a rise in amino acids and a fall in protein. Potassium deficiency also led to a rise in amino acids, while protein remained unchanged. It might be noted in this connection that the authors' con- clusion that increased amino acids with potassium deficiency are the Robert A. Steinberg 383 result of increased hydrolysis and not hindrance in synthesis of protein is not supported by their data. Wall (59) discussed this point also and disagreed with their interpretation particularly because of the presence of ample sugar in the leaves to prevent hydrolysis. Another interesting comparison of these data is one based on the hypothesis that mineral deficiency leads to a disproportionate content of amino acids and reducing sugar as compared with normal plants. Computation of the average deficiency-control ratios for all chemical elements discloses that in 9 out of 10 instances amino nitrogen was relatively greater than protein nitrogen in leaves, and only 4 out of 8 times in the stems. The analogous figures for reducing sugar and carbo- hydrate were 5 out of 8 in leaves, and 3 out of 8 in the stems. Since symptoms of abnormal nutrition are primarily localized in the leaves, we may assume that mineral deficiencies lead to an accumulation of amino acids as compared to protein, and of reducing sugar as com- pared to carbohydrate. Extent and duration of deficiencies will probably be found to cause progressive and opposite reactions in proportions of carbohydrates. An accumulation of nitrate in deficient plants appears to be a general phenomenon of mineral deficiency excepting nitrogen and boron. Since deficiency symptoms appear in plants supplied with ammonium nitro- gen, it must be assumed that inability to utilize nitrate is a result and not a cause of symptoms. With the fungus, Aspergillus niger, only in the case of molybdenum has a mineral deficiency been found dependent on the source of nitrogen (45). Further studies with green plants will presumably lead to similar results, inasmuch as Mulder (28) was able to demonstrate a diminution effect of ammonium nitrogen on molyb- denum deficiency in tomatoes, barley, and oats. In the tung and other trees, however, nitrate reduction apparently occurs in the roots instead of the leaves, and no evidence has as yet been found that mineral deficiency can cause an accumulation of nitrate. Nitrates, it is known, can be caused to accumulate in leaves of herbaceous plants by manipulation of climatic conditions irrespective of mineral deficiencies. It is evident, therefore, that caution should be exercised in attributing the accumulation of an intake material or metabolite primarily to mineral nutrition. Such an accumulation may 384 Mineral Nutrition of Plants be due instead to a diminution in demand due to a decreased rate of growth. The experimental results here summarized have been obtained almost wholly with herbaceous plants; the pineapple, guayule, and tung tree are the only exceptions. Analytical results for the tung tree disagreed in many respects with those for herbaceous plants. A decision on whether this disagreement is caused by an intrinsic difference in metabolism between these two types of vegetation must await the accumulation of data on a large variety of trees. SUMMARY Mineral deficiencies in plants appear to lead to a relative accumula- tion of amino acids and reducing sugars. Protein may or may not de- crease depending on the extent and duration of deficiencies. Accumula- tion of nitrates appears to be a result of diminished growth with mineral deficiency, except in the case of molybdenum. The alterations in propor- tions of nitrogen fractions lead to the assumption that all minerals participate directly or indirectly in nitrogen metabolism. Carbohydrates may temporarily accumulate during the early stages of deficiency. Evidence exists for a localization effect in that plant parts show a de- gree of parallelism between severity of visual and chemical symptoms. It appears useless to attempt to draw any connection between indi- vidual deficiencies and physiological responses in the plant, since pro- tein metabolism is a basic reaction for all. REFERENCES 1. Bennett, J. P., Minor Elements. Evidence and Concepts on Functions, Deficiencies, and Excesses. Ed. by Firman E. Bear and Herminie Broedel Kitchen. (Baltimore, Md., 1945). 2. Breon, W. S., and Gillam, W. S., Plant Physiol., 19:649 (1944). 3. Briggs, George B., Plant Physiol., 18:415 (1943). 4. Burrell, R. C., Hotan. Gaz., 82:320 (1926). 5. Calvin, M., and Benson, A. A., Science, 107:476 (1948). 6. Chibnall, Albert Charles, Protein Metabolism in the Plant (New Haven, Yale Univ. Press, 1939). 7. Cooil, Bruce J., and Slattery, M. C, Plant Physiol., 23:425 (1948). 8. Drosdoef, Matthew, Sell, Harold M., and Gilbert, Seymour G., Plant Physiol., 22:538 (1947). Robert A. Steinberg 385 9. Eaton, Scott V., Botan. Gaz., 97:68 (1935). 10. , Botan. Gaz., 102:536 (1941). 11. Eckerson, Sophia H., Contribs. Boyce Thompson Inst., 3:197 (1931). 12. Gilbert, Seymour G., Sell, Harold M., and Drosdoff, Matthew, Plant Physiol., 21:290 (1946). 13. Gregory, F. G., and Baptiste, E. C. D., Ann. Botany, 50:579 (1936). 14. Gregory, F. G., and Sen, P. K., Ann. Botany, N.S., 1:521 (1937). 15. Haas, A. R. C., Botan. Gaz., 97:794 (1936). 16. Hartt, Constance Endicott, Plant Physiol., 9:453 (1934). 17. Hibbard, H. P., and Grigsby, B. H., Mich. Expt. Sta. Tech. Bull. 141. 39 PP- O93-0- 18. Janssen, George, and Bartholomew, R. P., /. Am. Soc. Agron., 24: 667 (1932). 19. Janssen, George, Bartholomew, R. P., and Watts, V. M., Ar\. Agr. Expt. Sta. Bull. 310. 43 pp. (1934). 20. Johnston, Earl S., and Dore, W. H., Plant Physiol, 4:31 (1929)- 21. Jones, Winston M., Plant Physiol., 11:565 (1936). 22. Kraybill, H. R., Ind. Eng. Chem., 22:275 (1930). 23. , and Smith, T. O., N. H. Agr. Expt. Sta. Bull. 212: 7-9 (1924). 24. Lucas, Robert E., Soil Sci., 65:461 (1948). 25. MacGillivray, John H., /. Agr. Research, 34:97 (1927). 26. MacVicar, Robert, and Burris, R. H., /. Biol. Chem. 176:511 (1948). 27. Mothes, K., Planta, 1:472 (1926). 28. Mulder, E. G., Plant and Soil, 1:94 (1948). 29. Nightingale, Gordon T., Botan. Rev., 133:85 (1937). 30. , Botan. Rev., 14:185 (1948). 31. , Addoms, R. M., Robbins, W. R., and Schermerhorn, L. G., Plant Physiol., 6:605 (1931)- 32. , Schermerhorn, L. G., and Robbins, W. R., N. /. Agr. Expt. Sta. Bull. 499 (1930). 33. , Schermerhorn, L. G., and Robbins, W. R., Plant Physiol., 7:565 (1932). 34. Phillips, T. G.. Smith, T. O., and Dearborn, R. B., N. H. Agr. Expt. Sta. Tech. Bull, ^g (1934). 35. Phillips, T. G., Smith, T. O., and Helper, J. R., N. H. Agr. Expt. Sta. Tech. Bull. j$ (1939). 36. Richards, F. J., and Templeman, W. G., Ann. Botany, 50:367 (1936). 37. Schmalfuss, Karl, Phytopath. Z. 5:207 (1932). 38. Scripture, P. N., and McHargue, J. S., /. Am. Soc. Agron., 35:988 (i943)- 39. Scripture, P. N., and McHargue, J. S.. /. Am. Soc. Agron., 36:864 (1944). 40. Sideris, C. P., and Young, H. Y., Plant Physiol., 19:52 (1944). 386 Mineral Nutrition of Plants 41. Sideris, C. P., and Young, H. Y., Plant Physiol., 20:649 (1945). 42. Sideris, C. P., and Young, H. Y., Plant Physiol., 21:218 (1946). 43. Skok, John, Plant Physiol., 16:145 (I941)- 44. Smirnov, A., Strom, E., and Kuznetzov, S., Bull. acad. sci. U.R.S.S., Ser. Biol 2:265 (1938). 45. Steinberg, Robert A., /. Agr. Research, 59:731 (1939). 46. , /. Agr. Research, 64:455 (1942). 47. , /. Agr. Research, 64:615 (1942). 48. , /. Agr. Research, 75:81 (1947). 49. , /• Agr. Research, 78:733 (1949). 50. — , Bowling, John D., and McMurtrey, J. E. Jr., Plant Physiol., 25:279 (1950). 51. Stepka, W., Benson, A. A., and Calvin, M., Science, 108:304 (1948). 52. Steward, F. C, and Street, H. E., Plant Physiol., 21:155 (1946). 53. Turtschin, Th. W., Z. Pfianzenernahr. Diingung u. Boden\., 44:65 (1936). 54. Vickery, Hubert Bradford, Cold Spring Harbor Symposia on Quanti- tative Biology, 6:67 (1938). 55. , Biological Symposia, 5:3 ( 1 94 1 ) . 56. , Pucher, George W., Schoenheimer, Rudolph, and Ritten- house, D. /. Biol. Chem., 135:531 (1940). 57. Viets, F. G., Jr., Moxon, A. L., and Whitehead, E. I., Plant Physiol., 21:271 (1946). 58. Wadleigh, Cecil H., and Shive, John W., Soil Sci., 47:33 (1939)- 59. Wall, Monroe E., Soil Sci., 47:143 (1939). 60. , Soil Sci., 49:393 (1940). 61. Wood, J. P., and Petrie, A. H. K., Australian J. Exptl. Biol. Med. Sci., 20:249 (1942). MODIFYING INFLUENCES OF VARIOUS ENVIRON MENTAL FACTORS UPON MINERAL NUTRITION CHAPTER I O Light as a Modifying In- fluence on the Mineral Nutrition of Plants ROBERT B. WITHROW L iight is not known to play any direct role indispensable for the absorption, movement, or metabolism of the mineral nutrients. The ions of the mineral macroelements in the inorganic state have no absorption bands making it possible for direct light interaction. A few of the microelements in the ionized state, such as iron, manganese, and copper, have absorption bands in the visible spectrum, but thus far no strong evidence has appeared that light absorption by such ions is of any physiological significance. The effect of light on mineral nutrition, therefore, must be chiefly an indirect one resulting from the tempera- ture rise concomitant with the absorption of light and from the basic photochemical reactions occurring in plants, such as photosynthesis, chlorophyll synthesis, photomorphogenesis, and photoperiodism. In many cases, the various mineral elements are essential for a photo- chemical process, either as a constituent of an enzyme system or as a component of a pigment or other substance involved in the photo- chemical process. In such cases, it is apparent that the rate of the photo- chemical process frequently will influence the mobilization and utili- zation of the essential mineral elements. In addition, light can affect the physiology of growth, of which mineral metabolism is a part, through various rather poorly understood photochemical effects such as alteration of permeability, changes in protoplasmic streaming, and various photochemical oxidations. The most conspicuous indirect role of light in mineral metabolism is through photosynthesis. A carbohydrate supply is necessary as a 39° Mineral Nutrition of Plants source of energy for the accumulation of mineral nutrients by roots and for carrying out all other chemical reactions driven energetically by respiratory reactions. The mineral elements enter into synthesis with carbohydrate-derived materials to form the metabolic and structural components of plants. It therefore follows that the mineral nutrient requirement of plants is directly dependent upon the carbohydrate supply and, in turn, upon photosynthesis. While there are many photochemical reactions in plants necessary for normal growth and development, photosynthesis is the principal energy-converting reaction and consequently requires relatively high light intensities to maintain optimum growth. So far as is now known, all the other photochemical reactions are saturated at relatively low light intensities of the order of 20 foot-candles or less. In the case of photosynthesis in higher plants, on the other hand, saturation does not occur until light intensities of several thousand foot-candles have been attained. In the case of wheat, Hoover et al. (22) have shown that the relation between the rate of photosynthesis and light intensity is very nearly a linear one up to about 1000 foot-candles at normal atmospheric carbon dioxide concentrations of 0.03 volume per cent. Above 1000-2000 foot-candles, photosynthesis increases with increase in light intensity, but less rapidly than at lower values. For the other photochemical processes such as chlorophyll synthesis, photoperiodism, and phototropism, the intensity of daylight between sunrise and sunset seldom becomes a factor critically limiting the rate of reaction. For photoperiodism and certain morphogenic effects, the daylight factor that becomes limiting is the duration of the lighted period. In view of these considerations, the relative length of day and night and incident solar energy become important in relation to mineral nu- trition. At 40 degrees latitude in the temperate zone, the natural day length between sunrise and sunset varies from about nine hours to a little over fifteen hours. The range is greater in the more northern lati- tudes and less in the southern regions. The average daily total solar and sky radiation on a horizontal surface at a 40-degrees latitude station (79), however, is 80 per cent less during the winter months than during June and July. Since the reduction in day length is only about 40 per Robert B. Withrow 391 cent, an increase in cloudy weather during the winter accounts for much of the decrease in available radiation. However, data compiled by Kimball {2j) have shown that the maximum intensity on a clear day in December is within 10 per cent of the highest intensities attained in June. This generalization applies to data for clear cloudless days from Alaska to Puerto Rico. Thus, greenhouse crops in the winter in the midwestern and northeastern portions of the United States are sub- jected to short days and relatively low daylight intensities, both of which greatly limit the nutrient requirements of the plants. THERMAL EFFECTS The absorption of sunlight by leaves and other portions of the shoot and by the soil surface frequently results in shoot and root tempera- tures significantly higher than the ambient air temperature. This rise in temperature due to the absorption of radiant energy may become sufficiently high to significantly accelerate thermal reactions and has been shown to alter markedly the transpiration rate; it undoubtedly affects enzymatic reactions and causes changes in form and rate of growth. When means of heat dissipation are restricted, lethal tempera- tures may be reached. Horizontal leaf surfaces, at midday during clear weather in the tem- perate zone, are exposed to sunlight intensities ranging from 1.2 to 1.5 gram calories per square centimeter, minute. And of this total incident energy, 20 to 30 per cent is reflected by thin leaves, a much smaller pro- portion is transmitted, and the remainder is absorbed. Most of the ab- sorbed energy is degraded to heat and manifests itself as a temperature rise of the exposed portions of the shoot, while only a small portion, less than 5 per cent, is used in photosynthesis. Curtis (9) has found that the temperature of citrus leaves in intense sunlight may be 10 to 15 degrees above the air temperature and that all leaves exposed to intense sunlight had a higher temperature than that of the surrounding air, regardless of wind velocity or relative humidity, although high wind velocities could reduce the temperature by a matter of 10 degrees. Miller and Saunders (^5) have reported a diurnal vari- ation in leaf temperature and transpiration rate of corn, both of which attain maxima coinciding with the maximum of sunlight intensity. 392 Mineral Nutrition of Plants The marked effect of temperature is shown by the fact that a 5-degree rise in temperature of the saturated intercellular spaces of the leaf has the same effect in accelerating evaporation as a 30 to 40 per cent fall in relative humidity (9). Increased transpiration usually results in some increased absorption of mineral nutrients. Freeland (12) (Table I), Wright (yj), and others have shown that the uptake of the macroelements is directly correlated with the transpiration rate, but is not necessarily proportional to the TABLE I Data on the Amount of Mineral Absorption in Plants with High Transpiration (H.T.) and Low Transpiration (L.T.) (from R. O. Freeland) Total Minerals Calcium Phosphorus Potassium Water Absorbed, Absorbed, Absorbed, Absorbed, Absorbed, g- g- g- g- cc. Corn H.T L.T. 2-75 2.54 0. 17 0.18 0.25 0.15 0. 16 0. 12 1625 645 Bean H.T 1.86 0. 17 0.18 0.25 1175 L.T 1 .42 0.13 0.15 0.03 425 rate of water uptake. Broyer and Hoagland (4) have emphasized that the metabolic condition of the plant is a very important factor in de- termining the influence of transpiration on the uptake of salts by roots. The salt uptake of young barley plants having an initially low salt, high sugar composition was found to be principally dependent on aeration and temperature, with the data showing only a slight increase in salt uptake when the transpiration rate and water absorption were increased by light and low relative humidity (Table II). High salt and low sugar plants, on the other hand, took up potassium and bromium ions at a considerably higher rate as the rate of transpiration and water absorp- tion was increased. These data indicate that water uptake by roots and salt absorption are relatively independent processes, but, under internal root conditions which are unfavorable for salt uptake (as in high salt, low sugar plants), an increased transpiration stream contributes ma- terially to increased salt uptake. Robert B. Whhrow 393 PERMEABILITY It has been known for many years that light in the visible and near- ultraviolet regions of the spectrum directly alters the capacity of many plant tissues to take up inorganic salts and organic solutes through processes other than those involving transpiration. Likewise, the ca- pacity to retain these solutes against osmotic gradients is influenced. Lepeschkin (29, jo) found that certain dyes are accumulated more rapidly in Elodea in light of relatively high intensities than in dark- TABLE II Influence of Transpiration on Absorption of Salt by Barley Plants of Low-Salt, High-Sugar, or High-Salt, Low-Sugar Status (from T. C. Broyer and D. R. Hoagland) Experimental Conditions Water Ab- sorbed, ml./g. fresh wt. Shoot Salt Absorbed in m.e. X io"/g. total fresh wt. (loss from culture) K Br Total Sugar, g./b Shoot Root High salt; low humidity, light 8.10 5.20 6 High salt; high humidity, light 2.58 3-24 4 High salt; high humidity, dark 1.49 1.59 2 Low salt; low humidity, light 9.60 10.85 9 Low salt; high humidity, light 3. 60 10.40 9 Low salt; high humidity, dark 2 . 52 8 . 75 9 07 24 15 52 65 *3 5 4 2.9 °-3 15 3 6.2 2.0 1 . 1 0.7 trace 3-5 2-5 0.8 ness. He concluded that the protoplasmic permeability was increased and that the most effective spectral region was in the blue and near- ultraviolet range from 320 to 420 mu. Similar results were obtained by Oflford and d'Urbal (42) with Nitella. Jacques (2j) observed that the uptake of ammonium ions from sea water as well as the exosmosis of ammonium ions to ammonium-free sea water is higher in Valonia plants exposed to daylight than those kept in the dark. On the other hand, Zycha (60) and Jacques and Osterhout (24) have found that the accumulation of potassium in Nitella was relatively independent of radiation. 394 Mineral Nutrition of Plants Recently, Lepeschkin (j/) has reported that sunlight and ultraviolet radiation from a mercury arc accelerates the exosmosis of salts from the leaves of Sambucus and Parthenocissus and from potato tuber tissue. These results were obtained by measuring the changes in electrical con- ductivity of distilled water in contact with the excised leaves and potato tuber discs. Hoagland, Hibbard, and Davis (21) and Lundegardh (■?■?) have shown that Nitella cells and excised roots, respectively, can carry out ion accumulation only in the presence of an adequate supply of oxi- dizable respiration substrates as carbohydrates. Thus, ion accumulation by green tissues is indirectly accelerated by light through photosynthesis. This factor seriously complicates the interpretation of experimental data obtained with green leaves. The fact that the shorter wave lengths of the visible and the near ultraviolet are the most effective spectral re- gions and that the effects are observable in nonchlorophyllous tissue, such as potato, would indicate that light appreciably alters the capacity of cells to exchange solutes other than by providing a source of carbo- hydrates for accumulation processes. Stalfelt (49) and Virgin (52), using centrifugation methods, meas- ured the change in viscosity in the leaves of Helodea densa in response to incandescent lamp irradiation. Virgin observed cyclic short- and long- term fluctuations, the pattern of which varied with light intensity; at intensities of 200 to 2000 foot-candles, a short transient increase in vis- cosity was observed followed by a rapid fall, while at the very low in- tensity of 0.05 foot-candle the viscosity decreased and remained low for several hours of continuous irradiation. The significance of these results in terms of the movement of solutes is not immediately apparent, but experiments of this nature should be borne in mind in evaluating the effect of light on solute movement and metabolism. It is possible that light-induced permeability and viscosity changes may be the result of sensitized photochemical oxidations similar to those recently observed by Galston (14, 75) on the photooxidation of auxin and other cellular constituents with riboflavin, a fluorescing pig- ment of the respiratory enzyme systems, as the photosensitizer. Irradi- ation within the range of the principal absorption band of riboflavin, about 4600 A., results in the transfer of the energy to neighboring mole- Robert B. Wi throw 395 cules which may then be oxidized in the presence of molecular oxygen. Several amino acids and indoleacetic acid have been shown by Galston to be readily photooxidized by riboflavin. One interesting feature about such a sensitized photooxidation is that the photosensitizing pigment is not itself rapidly destroyed in the process. It merely absorbs a quan- tum of light, becomes activated with excess energy, and transfers the energy to a neighboring molecule as activation energy or re-emits the energy as fluorescence, depending upon the opportunity for suitable collision. The activated neighboring molecule then uses the energy for reaction with molecular oxygen or degrades it to thermal energy. It is possible that riboflavin and other fluorescing pigments may carry out a wide variety of photochemical oxidations which may account for many of the rather obscure effects of light on plant cells such as light-modified permeability, viscosity, and protoplasmic streaming. The recent work of Goodwin and Kavanagh (iy) on fluorescing substances in roots in- dicates that plant cells may contain a number of fluorescent substances capable of entering into such photooxidation systems. It is impossible at the present time to evaluate the over-all physiologi- cal significance of light effects on permeability, photooxidations, and other such processes. Further investigations may show that these play a more significant part in the metabolic activities of the shoot than is now assigned to them. As far as roots growing in soil are concerned, however, these phenomena are obviously unimportant as directly con- trolling processes but they may exercise definite indirect effects. PHOTOPERIODISM Photoperiodism involves photochemical and thermal reactions initi- ated in the leaf which through the translocation of one or more hor- mones control the course of development of the terminal shoot meri- stems and in turn markedly influence the rate of shoot growth in many species. Higher plants fall into four classes as regards their flowering response to photoperiod on normal daily cycles: (a) long-day plants which flower on photoperiods longer than a certain critical day length, but remain vegetative on shorter photoperiods; (b) short-day plants which flower on photoperiods shorter than a certain critical clay length; (c) intermediate plants which flower only on photoperiods within a 396 Mineral Nutrition of Plants narrow photoperiod range; and (d) day-neutral plants indifferent to photoperiod as regards flower bud initiation. In all these classes are plants whose flowering is also influenced by temperature. With some of these types in certain combinations of day and night temperatures, photoperiod has little influence on the course of development, the determination of the initiation of flower buds being controlled chiefly by temperature. Numerous investigations have been conducted in an endeavor to cor- relate photoperiodic reactions with mineral nutrient supply and metab- olism. With but few reported exceptions, variations in the mineral nutrient supply within the range capable of supporting growth have had no determinative effect capable of altering the course of floral initiation. The two environmental factors which appear to be deter- minative are relative length of the light and dark periods and tempera- ture. In spite of the negative results obtained, however, several inter- esting correlations have been reported on the influence of mineral nutrient levels on the rate of appearance of floral primordia and the effect of photoperiod on the rate of appearance of nutrient deficiency symptoms. Nitrogen relationships Particular attention has been paid to nitrogen relationships. While the level of nitrogen appears to play no critical role in determining the induction of flowering, low levels of nitrogen do significantly affect the time of appearance of macroscopic flower buds. In short photo- periods, decreasing or removing the nitrogen supply decreases the rate of growth very markedly. In such cases, it has been found that the macroscopic flower buds of short-day plants often appear later than those of plants growing in a high nitrogen medium. In a long photo- period such plants are vegetative at all nitrogen levels. This situa- tion is exemplified in short-day plants, such as chrysanthemum (7), Xanthium (jo, 5S), soybean (46, 58), Kalanchoe (7), Tinantia fitgax (7), Setaria italica (7), Tithonia speciosa (5S), and Salvia splendens (58). For short-day plants in general, abundant nitrogen supply appears to hasten flowering. Without an adequate nitrogen supply, the number of floral buds set is sparse and fruiting often fails to occur. Robert B. With ro w 397 Conversely, certain long-day plants in a long photoperiod favorable for flowering and deficient in nitrogen tend to develop macroscopic flower buds earlier than high nitrogen plants. This has been found to be true with barley (j), hard wheat (7), Iberis (7), spinach (28, 5$), and lettuce (7). In a long photoperiod, the rate of growth is not de- creased to the same extent by decreasing the nitrogen supply as it is in the case of a short photoperiod (5#). In this same relation, Scully et al. (45) found that in onion, which is a long-day type of plant and forms bulbs only in a long day, a low nitrogen level tends to accelerate the rate of development of bulbing. These observations do not sub- stantiate the oft-repeated horticultural concept that a low nitrogen level hastens maturity and the flowering of plants. This correlation appears to exist only for certain of the long-day and possibly the day- neutral types of plants. B oron The rate of appearance and severity of several of the micronutrient deficiency symptoms have been found to have a definite relation to photoperiod. Warington (55) observed that plants growing on a boron- deficient medium during the spring and fall months required a longer time for the development of the typical deficiency symptoms than similar plants grown during the summer. She investigated the possible relationships between temperature and day length to these observations and concluded that it was a photoperiod-controlled reaction. With broadbean, scarlet runner bean, Mandarin soybean, and barley, a nine- hour day very definitely retarded the appearance and progress of the boron deficiency symptoms as contrasted with a normal day length of over 15 hours. Biloxi soybean and garden pea exhibited no delay in appearance of symptoms in relation to photoperiod. The deficiency symptoms characteristic of lack of boron were similar regardless of time of appearance. MacVicar and Struckmeyer (34) further demonstrated that Xan- thium, tomato, sunflower, and buckwheat exhibited the usual boron deficiency symptoms in a boron-deficient medium during a 16-hour photoperiod. In a short photoperiod of nine hours, the deficiency symp- toms did not appear in these plants. These experiments were conducted 39$ Mineral Nutrition of Plants during the winter months when the daylight intensity was low and the photoperiod was short. Supplemental incandescent lamp irradiation was used to lengthen the photoperiod. Thus, the light intensity was relatively low as compared with summer conditions and the growth rate was correspondingly lower. These authors conclude that the lack of severity of boron deficiency symptoms under a short photoperiod is correlated with cessation of cambial activity and the occurrence of floral induction. Manganese and iron Steckel (50), working with soybean, has observed the appearance of deficiency symptoms for iron and manganese under conditions of different light intensities and day lengths, using varying ratios of iron and manganese. He found that the time of appearance of the symp- toms could be correlated with light conditions. The symptoms ap- peared earlier at high light intensities than at low and earlier in long photoperiods than in short. Weight and top-root ratio data indicated that one of the chief factors probably responsible for the delay in ap- pearance of the deficiency symptoms in short photoperiods and low intensities as compared with long photoperiods and high light inten- sities was a slower rate of growth. CHLOROPHYLL SYNTHESIS Chlorosis is one of the most conspicuous symptoms of both macro- nutrient and micronutrient deficiencies in higher plants and yet rela- tively little is known as to the biochemistry of chlorophyll synthesis and the mechanisms by which the lack of each element produces its own characteristic pattern of chlorophyll insufficiency. Nitrogen and magnesium are essential in the structure of the protochlorophyll and chlorophyll molecules. Galston (/j) has reported that the chloroplasts contain from 30 to 40 per cent of the total nitrogen in both chlorotic and green leaves of oat. Granick (18) found similar results for tomato and tobacco, with 10 per cent of the chloroplast nitrogen present in the chlorophyll molecule. About 10 per cent of the magnesium in the plant is found in chlorophyll, with magnesium comprising about 2.7 per cent of the chlorophyll molecule. Several of the micronutrients, such as iron, Robert B. Withrow 399 zinc, and copper, appear to participate in certain of the enzymatic steps of chlorophyll synthesis. Phosphorus may very likely enter into energy exchanges associated with the synthesis. The role of the other elements in causing chlorosis is obscure. In higher plants, chlorophyll synthesis is a photochemical reaction which proceeds at its maximum rate under relatively low light intensi- ties. It appears to be a photocatalytic type of reaction in which the light energy is absorbed by a pigment which presumably is protochlorophyll. Protochlorophyll is present in the dark-grown leaf in relatively low concentrations as compared with the ultimate maximum concentration attained for chlorophyll in the green leaf (47). During the synthesis of chlorophyll, therefore, there must be a concomitant synthesis of proto- chlorophyll. The synthesis of protochlorophyll appears to be a thermal reaction and not directly dependent upon light. In the algae, such as Chlorella, the synthesis of chlorophyll can be thermally activated and will proceed in the dark in the presence of sugars. Both protochlorophyll and chlorophyll are magnesium porphyrins which may be extracted with ether and other lipoid solvents. J. H. C. Smith (48) has recently initiated a program of study of chlorophyll synthesis and has redetermined the absorption spectrum of protochlorophyll. It contains two strong maxima which, in ether extracts, appear at 623 and 432 mu, with the blue maximum much stronger than the red. There is appreciable absorption, however, throughout the ultraviolet and visible range from 250 to 650 mu. The maxima coincide quite closely with the position of the action spectrum maxima of chlorophyll synthesis determined with light filters by S. Frank (//). Smith (48) reported that, in the synthesis of chlorophyll, there is a concomitant decrease in protochlorophyll concentration and an increase in ether-soluble magnesium and phosphorus, the time-course patterns of which coincide closely with the appearance of chlorophyll. In the excised leaf, detached from the root system, the total ash, including magnesium and phosphorus, is relatively unchanged, but there is a large increase in ether-soluble fractions of both these elements. For leaves on an intact plant, however, there is an increase in ash corre- sponding with increases in lipoidal magnesium and phosphorus (Figure 400 Mineral Nutrition of Plants i). Thus, there appears to be a mobilization of magnesium and phos- phorus into leaves of greening barley seedlings and an increase in ether- soluble organic materials. Smith concludes that, in view of the relatively low concentration of protochlorophyll attained under equilibrium con- ditions in the dark, chlorophyll synthesis must be a rather complex series of reactions involving protochlorophyll synthesis with chlorophyll ETHER ASH SOLUBLE SUB 3S0 7000 300 6000 eso 5000 200 4000 ISO 3O00 IOO 3000 so I0OO Chlorophyll x/0{ Et sol. Material x!0i G Ash, xlO' Ma - Chi a 10 7 P-Et sol. rIO7 30 to SO 60 IRRADIATION TIME -HOURS. Figure i. Effect of light on chlorophyll and other associated constituents of dark-grown barley leaves on intact plants. — }. H. C. Smith (47). Composition given for chlorophyll, total ether-soluble material, chloro- phyll magnesium, ether-soluble phosphorus, and total ash. synthesis and the mobilization of magnesium and phosphorus into the synthesizing system. The significance of the accompanying increase of ether-extractable phosphorus is not clear, but phosphorylations may be involved. This increase is in harmony with Stoklasa's (5/) original discovery that seedlings germinating in the light contain more ether- soluble and alcohol-soluble phosphorus than dark-grown plants. These results on the magnesium and phosphorus mobilization and conversion into ether-soluble compounds during chlorophyll synthesis Robert B. Withrow 401 may explain, in part at least, observations that magnesium and phos- phorus are frequently taken up at parallel rates in higher plants. Iron has long been recognized as an essential micronutrient for the synthesis of chlorophyll. Jacobson (25) determined the iron content of pear, tobacco, and corn leaves in relation to varying degrees of iron chlorosis. He found that both total iron and acid-soluble iron paralleled the chlorophyll content, but he found also that some of the iron is inac- tive in chlorophyll synthesis and this must be exceeded before chloro- phyll formation can proceed. He observed that the "active" or ferrous iron was localized chiefly in the chloroplasts. This is in harmony with the idea advanced by Mommaerts (36) that chlorophyll in the leaf exists in combination with a protein and that such a chlorophyll-protein com- plex contains iron. In addition iron-containing compounds of this type have been isolated from corn and tobacco. It has been suggested that the iron is associated closely with a protein and that this protein complex, which apparently catalyzes chlorophyll formation, serves as a "chloro- phyll enzyme." Neish (40) has found that there is a greater percentage of copper in the chloroplasts than in the remainder of the leaf. In Birfolium pratens, the proportion was 74.6 per cent of the total. It has been found in some plants that there is a greater concentration and longer persistence of chlorophyll in copper-treated plants. Working with a variety of plants, Okunsov (43) found a marked increase in chlorophyll content of the leaves on the addition of copper sulfate spray when the plants were grown on a copper-deficient turf. The increased chlorophyll was held as principally due to a retarding effect of copper on the decomposition of chlorophyll with age. Arnon (/) has shown copper to be a com- ponent of the chloroplast of chard as a polyphenol oxidase. This is a copper-protein complex which may play a role in the evolution of oxygen by isolated chloroplasts during the water-splitting phase of photosynthesis. PHOTOSYNTHESIS Current interpretation of photosynthesis as an over-all process is tend- ing increasingly toward the view that the reduction of carbon dioxide and the fission of water with the evolution of oxygen is a relatively 402 Mineral Nutrition of Plants complicated, many-step process closely integrated with the metabolic activities of the chlorophyllous cell. Of particular interest to the subject of this paper are evidences that phosphorus turnover and nitrate reduc- tion are part of the photosynthetic process. Before carbon dioxide enters the reduction phases of photosynthesis, it is fixed in the cytoplasm outside the chloroplast. The fixation is a nonphotochemical, enzymatic carboxylation in which organic acids are formed. This first step, known as carbon dioxide fixation, is not the characteristic feature of photosynthesis nor is it confined to photo- synthesizing cells as it occurs in both autotrophic and heterotrophic organisms. After fixation, the carboxylated product is carried through the various stages of reduction; a hexose derivative usually appears as the end product. The most characteristic feature of photosynthesis is the photochemi- cal fission of water with the evolution of oxygen. It has been shown recently that isolated chloroplasts are capable of carrying out this stage of the reaction as a photochemical process. All the steps, from the initial fixation to the final fission of water and reduction of carbon dioxide, involve energy transfers which must require the storage of energy for very short periods in an available form in order to prevent back reac- tions and allow time for the next steps. Phosphorus The animal biochemists have shown that the energy transformation in the conversion of chemical energy to mechanical energy in muscle and the conversion of chemical energy from one form to another in the processes of respiration occur through the agency of phosphorylated compounds containing high-energy phosphate bonds. Several investiga- tors have presumed that some of the energy storage processes of photo- synthesis may involve similar energy transformations, postulating that inorganic phosphorus would flow into the photosynthesizing system when energy is stored but not used, and be released during energy utilization phases. The downgrade flow of energy through respiratory processes has been shown to be carried out in plants, animals, and microorganisms by way of the high-energy phosphate bonds occurring in phosphorylated Robert B.Withrow 403 compounds such as adenosine triphosphate. Kalckar (26) and Lipmann (52) have very thoroughly reviewed the biochemistry of this subject. The role of phosphorus in the upgrade flow of energy in such organisms as the chemautotrophic bacteria which reduce carbon dioxide by the oxidation of reduced compounds as sulfur and hydrogen and in the photosynthetic organisms has been investigated only recently. Vogler and his co-workers (5^) have studied the exchange of phos- phorus between cells of Thiobacilhis thiooxidans and their environ- ment in relation to the oxidation of sulfur and the utilization of car- bon dioxide. This organism is a sulfur bacterium which synthesizes the materials needed for cell growth by the fixation and reduction of carbon dioxide with energy derived from sulfur oxidation. In the absence of carbon dioxide, Vogler found that sulfur can be oxidized and the energy stored within the cell and used later for the fixation of carbon dioxide under conditions unfavorable for sulfur oxidation. It was not determined experimentally whether the energy stored in this case could also be used for the reduction of carbon dioxide. It was sug- gested that, during sulfur oxidation in the absence of carbon dioxide, some of the energy is stored as phosphate bond energy and later used for carbon dioxide fixation with the release of inorganic phosphate. During sulfur oxidation it was observed that, in the absence of carbon dioxide, inorganic phosphate was taken up from the external medium. Subsequently, in the absence of sulfur oxidation and in the presence of carbon dioxide, some phosphate was released to the external medium, presumably by the breakdown of energy-rich phosphate com- pounds synthesized during the sulfur oxidation phase. However, dur- ing normal sulfur oxidation in the presence of an adequate supply of carbon dioxide, no appreciable exchange of phosphate occurred, as phosphorylation and dephosphorylation presumably were in equi- librium. Adenosine triphosphate was identified as one of the phosphate compounds formed during the oxidation of sulfur and is possibly the phosphorylated storage product. Vogler suggested that in photosynthe- sis, energy from light absorption may be stored in a similar manner in high-energy phosphate bonds of phosphorylated compounds and later released for the fixation of carbon dioxide. Ruben (44) measured the free energy of carbon dioxide fixation in 404 Mineral Nutrition of Plants Chlorella and found that it was a very rapid reaction, even at low partial pressures of atmospheric carbon dioxide, indicating a low free energy per mol of about -2 kilogram-calories. The free energy of reac- tion of known carboxylations involving aliphatic and aromatic com- pounds is very much higher, being of the order of 10 kilogram-calories. Ruben postulated that high-energy phosphate bonds which have a free energy of 8 to 12 kilogram-calories per bond might be involved in this reaction. Emerson, Stauffer, and Umbreit (10) and Ochoa (41) have postu- lated that the later steps following fixation of the carbon dioxide and involving the transfer of energy of the absorbed light quanta probably also occur through the agency of high-energy phosphate bonds. In endeavoring to test these hypotheses, Aronoff and Calvin (2) were unable to show any significant correlation between phosphate turnover and photosynthesis in Chlorella using P32 as a tracer. Gest and Kamen (/6), however, also used P32 in Scenedesmus and Chlorella and found that such turnover studies could be carried out with consistent results only when low phosphate cells were used. In such cells, the soluble phosphate, readily exchangeable during washing processes, is reduced to a minimum and errors resulting from such manipulations are con- siderably reduced. They found that phosphate turnover in this case could be correlated directly with photosynthetic activity. In Figure 2 are plotted the effects of light intensity on the trichlorace- tic acid-insoluble phosphate in Scenedesmus. The cells were grown on a low phosphate medium and the trichloracetic acid-insoluble fraction was found to be relatively stable as far as leaching was concerned. These data show that light markedly accelerates the formation of insoluble phosphate by the cells. Cyanide in Chlorella depressed the phosphate uptake; since the "ground respiration" of plant cells is cyanide resistant, these results indicate that the accelerated uptake is not correlated with respiratory activity that results directly from in- creased photosynthates. These experiments indicate that ester phos- phate may be formed as a result of light absorption, but the evidence is insufficient to determine whether such esterification is directly coupled with the reactions of photosynthesis. Wassink et al. {56, 57) have obtained somewhat similar results with Robert B. Withrow 405 photosynthesis in the purple sulfur bacterium, Chromatium, strain D. Irradiation caused a marked uptake of phosphate in an atmosphere of nitrogen and oxygen. When carbon dioxide was present, the uptake was less, indicating the possible utilization of ester phosphate in carbon di- oxide reduction. When the cells were transferred to darkness, there was a small release of inorganic phosphate, apparently associated with meta- bolic energy release. LIGHT + KCN ^-DAIiK +- KCN 60 120 180 TIME (MINUTES) Figure 2. Effect of light on P uptake by Scenedesmus D3. — Gest and Kamen ( 16). Data are for trichloroacetic acid-insoluble cellular residue. These few rather recent investigations strongly suggest that one of the most important roles of phosphorus in plants may be in phosphory- lation reactions involved in the upgrade flow of energy in photosyn- thesis. Nitrogen The two principal sources of inorganic nitrogen for the growth of higher plants are ammonium and nitrate. Ammonium nitrogen is already in a reduced form and in the dark is readily assimilated into organic forms of nitrogen. Nitrate nitrogen, on the other hand, must be reduced at least as far as nitrite before it can be assimilated into organic nitrogen. In the roots of most woody plants and legumes, this process takes place as a purely thermal reaction in the dark. Since the 406 Mineral Nutrition of Plants reaction is exergonic, energy is required which, in the case of nitrate reduction in roots, is derived from respiration. In 1920, Warburg and Negelein (54) observed a low assimilatory quotient with Chlorella which they assumed was due to the reduction of nitrates by some photosynthetic process. Burstrom (5, 6) later exten- sively studied nitrate reduction in the wheat plant. He found that the root was capable only of thermal reduction, which was independent of NITRATE HEXOSE p. MOL >* MO1- 60 300 40 200 - 20 100 OTHER ASSIMILATES •20 100 2000 FOOT CANDLES 3000 Figure 3. Nitrate reduction in excised wheat leaves as a function of photosynthetic rate over a 24-hour period in an atmosphere of about 10 per cent higher than normal air carbon dioxide. — H. Burstrom (5). Data are for apparent carbon dioxide assimilation, sugar formed, nitrate lost from leaf tissue, and other assimilates (CN compounds) formed. light, but required an external supply of manganese. Kinetical stuJies indicated that the reduction occurred at the very outer surface of the root cells and that, while manganese was essential, iron could partially substitute for manganese. In contrast to wheat roots, the leaves were completely unable to carry out nitrate reduction in the dark. In the presence of light and carbon dioxide, however, nitrate reduction occurred and the rate paral- leled very closely the photosynthetic rate as shown in Figure 3. This process does not appear to be the result of the synthesis of carbohydrates Robert B. Withrow 407 as such, since it will be noted that even when the light intensity drops below the compensation point and respiration is consuming more car- bohydrates than are being formed, nitrate reduction still proceeds. Neither does the process appear to be a simple photochemical reduction of nitrate as a substitute for carbonate since carbon dioxide is likewise essential as exhibited in Figure 4, where nitrate reduction was carried out in an atmosphere containing only about 10 per cent normal atmos- HEXOSE NITRATE /IMOL 40 20 -20 ■40 OTHER ASSIMILATES 1000 2000 FOOT CANDLES 3000 Figure 4. Nitrate reduction in excised wheat leaves in a low carbon dioxide atmosphere, about 10 per cent of normal air. — H. Burstrom (5). pheric carbon dioxide concentration. When the carbon dioxide con- centration is very low, nitrate reduction appears to take precedence over sugar formation; possibly some intermediate product of carbon dioxide reduction is utilized in the reduction of nitrate, thus blocking sugar formation. At high light intensities, sugar formation practically ceases in a low carbon dioxide atmosphere. These experiments at low carbon 408 Mineral Nutrition of Plants dioxide concentrations show that about 15 mols of carbon dioxide are used in the reduction of one mol of nitrate. Thus it appears that in the process of reduction of nitrate and carbon dioxide, some nitrogenous organic compound is formed. Myers (jj, 38) and Cramer and Myers (8) have studied nitrate reduction in Chlorella. This alga behaves like both the root and the leaf of wheat in that it is capable of both thermal and photochemical reduction of nitrate. The data in Table III show that the assimilatory TABLE III C02/02 Quotient for Chlorella Cells Grown at 40 Foot-candles and Studied at 40 Foot-candles (from Jack Myers, ref. jS) C02/02 by C02/0, Nitrogen Manometric Calculated from Source Measurement Cell Analysis NOs~ -0.68 -0.69 NH4 —0.94 —0.91 N03"+NH4+ -0.94 quotient (C02/02) is considerably less than one (0.68) in the presence of nitrate. In the presence of a mixture of nitrate and ammonium or ammonium alone, the assimilatory quotient is nearly one (0.94). When ammonium is present, photochemical nitrate reduction ceases. Some have objected to any concept which requires two such diverse pathways for nitrate reduction. This does not, however, appear to be a serious factor considering that in carbon dioxide reduction, two pathways have already been shown to exist in some of the lower organisms, one thermal and one photochemical. In the autotrophic chemosynthetic organisms, energy is derived from the oxidation of sulfur or hydrogen, which in turn is used for the fixation and reduction of carbon dioxide. In the presence of light, however, some of these organisms carry out a photosynthetic reduction instead of a thermal reduction. REFERENCES 1. Arnon, D., Plant Physiol., 24:1 (1949). 2. Aronoff, S., and Calvin, M., Plant Physiol., 23:351 (1948). Robert B. Withrow 409 3. Borodin, Irene, Bull. Applied Botany, Genetics, and Plant Breeding, (Leningrad ), 27:171 (1931). 4. Broyer, T. C, and Hoagland, D. R., Am. J. Botany, 30:261 (1943). 5. Burstrom, Hans, Ann. Roy. Agr. Coll. Sweden, 11:1 (1943). 6. -, Ann. Roy. Agr. Coll. Sweden, 13:1 (1946). 7. Cajlachjan, M. Ch., Compt. rend. acad. sci. U.R.S.S., 43:387 (1944). 8. Cramer, Marian, and Myers, J., /. Gen. Physiol., 32:93 (1948). 9. Curtis, O. F., Plant Physiol., 11:343 (1936). 10. Emerson, R. L., Stauffer, J. F., and Umbreit, W. W., Am. f. Botany, 31:107 (1944). 11. Frank, Sylvia R., /. Gen. Physiol., 29:157 (1946). 12. Freeland, R. O., Am. J. Botany, 24:373 (1937). 13. Galston, Arthur W., Am. J. Botany, 30:331 (1943). 14. , Proc. Nat. Acad. Sci. U.S., 35:10 (1949). 15. , and Baker, R. S., Science, 109:485 (1949). 16. Gest, Howard, and Kamen, Martin D., /. Biol. Chem., 176:299 (1948). 17. Goodwin, R. H., and Kavanagh, F., Bull. Torrey Botan. Club, 75:1 (1948). 18. Granick, Sam, Am. J. Botany, 25:561 (1938). 19. Hand, Irving F., Monthly Weather Rev., 65:415 (1937). 20. Hill, R., Nature, 139:881 (1937). 21. Hoagland, D. R., Hibbard, P. L., and Davis, A. R., /. Gen. Physiol., 10:21 (1926). 22. Hoover, W. H., Johnston, E. S., and Brackett, F. S., Smithsonian Misc. Coll. 8y:i (1934). 23. Jacques, A. G., /. Gen. Physiol., 22:501 (1939). 24. , and Osterhout, W. J. V., /. Gen. Physiol., 18:967 (1935). 25. Jacobson, L., Plant Physiol., 20:233 (1945). 26. Kalckar, H. M., Chem. Revs., 28:71 (1941). 27. Kimball, H. H., Monthly Weather Rev., 63:1 (1935). 28. Knott, J. E., Plant Physiol., 15:146 (1940). 29. Lepeschkin, W. W., Am. f. Botany, 17:953 (1930). 30. , Am. J. Botany, 19:547 (1932). 31. , Am. /. Botany, 35:254 (1948). 32. Lipmann, Fritz, Ann. Rev. Biochem., 12:1 (1943). 33. Lundegardh, H., Ann. Roy. Agr. Coll. Sweden, 16:372 (1949). 34. MacVicar, R., and Struckmeyer, B. E., Botan. Gaz., 107:454 (1946). 35. Miller, E. C, and Saunders, A. R., /. Agr. Research, 26:15 (1923). 36. Mommaerts, W. F. H. M., Magyar Biol. Kutatointezet Munl^dl, 15: 468 (1943)- 37. Myers, J., Plant Physiol., 22:590 (1947). 410 Mineral Nutrition of Plants 38. , in Franck, J., and Loomis, W. E., eds., Photosynthesis in Plants (Ames, Iowa, Iowa State College Press, 1949) 349. 39. Neidle, E. K., Botan. Gaz., 100:607 (1939). 40. Neish, A. C, Biochem. J., 33:293 (1939). 41. Ochoa, S., in Green, D. E., ed., Currents in Biochemical Research (Interscience Publishers, Inc., 1946) 165. 42. Offord, H. R., and d'Urbal, R. P., /. Agr. Research, 43:791 (1931). 43. Okunsov, M. M., Compt. rend. acad. sci. U.R.S.S., 54:829 (1946). 44. Ruben, S., /. Am. Chem. Soc, 65:279 (1943). 45. Scully, N. J., Parker, M. W., and Borthwick, H. A., Botan. Gaz., 107:52 (1945). 46. , Parker, M. W., and Borthwick, H. A., Botan. Gaz., 107:218 (i945)- 47. Smith, J. H. C, /. Am. Chem. Soc, 69:1492 (1947). 48. , Arch. Biochem., 19:449 (1948). 49. Stalfelt, M. G., Ar}(iv. /. Bot., 33A, No. 5:1 (1947). 50. Steckel, J., Unpublished data (1947). 51. Stoklasa, J., Ber., 29:2761 (1896). 52. Virgin, H., Physiologia Plantarum, 1:147 (1948). 53. Vogler, K. G., and Umbreit, W. W., /. Gen. Physiol., 26:157 (1943). 54. Warburg, O., and Negelein, E., Biochem. Z., 110:66 (1920). 55. Warington, K., Ann. Botany, 47:430 (1933). 56. Wassink, E. C., Enzymologia, 12:352 (1948). 57. , Tjia, J. E., and Wintermans, J. F. G. M., Proc. Konin\l. Nederland. A/{ad. Wetenschap., 52:412 (1949). 58. Withrow, Alice P., Butler Univ. Botan. Stud., 7:1 (1945). 59. Wright, K. E., Plant Physiol., 14:171 (1939). 60. Zycha, H., fahrb. wiss. Botan., 86:499 (1928). CHAPTER 1/ Soil Moisture and the Mineral Nutrition of Plants C. H. WADLEIGH and L. A. RICHARDS T, he mineral nutrition of plants is related in many ways, directly and indirectly, to soil moisture. It is the purpose of this paper to discuss some of these relations and it seems appropriate to introduce the subject with a review of some of the modes of behavior of water in soil. Particular attention will be given to the physical con- dition of moisture as it relates to the readiness with which roots may absorb moisture present at the soil-root interface, and also the readiness with which additional moisture at some distance from the roots can flow toward the roots to replace that which is absorbed. PHYSICAL CONDITION OF MOISTURE IN SOIL Soil particles are wetted by and have considerable attraction for water and, therefore, soils exhibit a capacity to take up and retain moisture. This moisture is distributed throughout the pore system of the soil and over the surface of the soil particles. Energy is involved in this adsorption because work is required to remove moisture from soil. The amount of work per unit amount of water removed will depend on the moisture content of the soil. This work can be expressed in terms of the pressure difference through which a unit volume of water must be transferred to effect removal of water from the soil. In saturated soil the pressure in the soil water approaches zero at atmospheric pres- sure. Water drops will form and drip from soil which is very wet, so the work that must be done against surface force action in order to effect water removal from very wet soil approaches zero. As soil dries 412 Mineral Nutrition of Plants out, more work must be done to extract water, and the equivalent negative pressure in the soil water increases. This pressure, as far as soil moisture energy relations are concerned, is equivalent to tensile stress and is referred to as soil moisture tension. Other factors being favorable, plants can grow and absorb moisture at soil moisture contents ranging from saturation down to some mini- mum moisture content that is associated with the wilting of plants and which depends on the texture of the soil. As this minimum moisture content is approached and the plant loses turgor, the rate of growth of the plant approaches zero. The wilting percentage may be as low as 2 or 3 per cent for sands or as high as 30 or 40 per cent for clay and peat soils. Normally, in the root zone in the field, soils do not become completely saturated. The soil moisture tension actually drops to zero only in the shallow surface layer. After heavy rain or irrigation, the water contained in the large pores of the soil drains away freely under the influence of gravity. This drainage rate is rapid at first, but after two or three days, in soils that do not have restricted drainage or a water table, the rate of drainage becomes negligible. The moisture content of the soil, after drainage becomes negligibly slow, is designated as the "field capacity." Consequently, it is common to refer to the moisture content of soil from field capacity to the wilting percentage as the "available range," or the range of moisture that is available for producing plant growth. The equivalent negative pressure in the soil ranges from zero in saturated soil to something of the order of 0.05 to 0.15 atmosphere at field capacity and on up to about 15 or 20 atmospheres at the wilting percentage for crop plants. This pressure scale is convenient since it is a direct measure of the work that plants must do to absorb water against surface force action in the soil. The relation between the mois- ture tension and the moisture content of soil is hyperbolic in nature (/oj) as illustrated by the curves in Figure 1. It is seen that a con- siderable amount of water may be withdrawn from wet soil before the tension rises more than an atmosphere or two. However, at tensions above 8 or 10 atmospheres a small decrease in moisture content cor- responds to a large increase in soil moisture tension. The curves in Figure 1 illustrate typical moisture-tension relations over the whole Wadleigh and Richards 413 plant growth moisture range and for a wide range of soil textures. The combination of a porous cup and a vacuum gage, which has been called a tensiometer, can be used for measuring soil moisture tension over the range from zero to about 0.85 atmosphere. The rela- tion of soil moisture content to soil moisture tension over the whole plant growth range has been studied by means of porous membrane 8 15 SOIL MOISTURE TENSION — ATMOSPHERES Figure i. Curves showing the relation between the soil moisture tension and the moisture content of the soil. apparatus {102). Recent improvements have been made in measuring methods so that it is now conveniently possible, by the freezing point procedure, to estimate soil moisture tension in nonsaline soils from freezing point depression values (104). Soil moisture tension is thus an experimentally usable scale for expressing the security or tenacity with which water is held in the soil by surface force action. If appreciable quantities of soluble salt are present in the soil, then an osmotic force also resists the absorption of water by plants and must also be taken into consideration. The sum of the soil moisture tension and the 414 Mineral Nutrition of Plants osmotic pressure of the soil solution has been called the total soil mois- ture stress (128) and can be used as a measure of the total amount of work that a plant must do to absorb a unit amount of water from the soil. The freezing point measurement directly gives an indication of the total soil moisture stress. Furr and Reeve (46) have shown that the rate of growth of sun- flower plants drops to zero as the soil moisture approaches the wilting percentage. There is increasing experimental evidence that the growth of crop plants ceases when the soil moisture stress rises into the range of 15 to 20 atmospheres (/jo). The course of events during moisture depletion in field soil may be set forth briefly by an illustration. Consider a newly planted hill of corn after a thorough soaking rain. When the rain stops, the moisture moves gradually out of the larger pores toward the water table which, for best crop growth, should be below the root zone. This water move- ment takes place primarily under the influence of gravitational force because, when the whole profile is very wet, gradients in soil moisture tension are small. Concomitant with the recession of water in the larger pores, air is drawn into these pores. As the soil approaches field capacity there is a reasonably continuous gaseous phase throughout the soil, but the mois- ture films covering the soil particles become so thin that additional moisture movement in these films is considerably restricted.* Thus, several days after the rain has stopped, the soil moisture ten- sion may rise into the range of 0.05 to 0.15 atmosphere (50 to 150 cm. of water) and moisture movement in the profile becomes considerably restricted. Water vapor loss will occur from the top few inches of sur- face soil due to air circulation and the elevated temperature of the surface soil, but moisture loss from soil layers a few inches below the surface is usually slow. When the corn seeds germinate and the plant roots develop in the neighborhood of the kernel, the rootlets remove moisture from the immediately adjacent soil and can build up soil * It should be kept in mind that the condition here referred to as field capacity can exist only in well-drained soils. When soil horizons having low permeability are present in the profile, or when a nearby ground water table exists, then the root zone may remain excessively wet for long periods. Wadleigh and Richards 415 moisture tensions of the order of 10 to 20 atmospheres. The high ten- sion in the soil moisture in the vicinity of the root sets up a force action in the soil-water system that tends to move water toward the root. This tendency of water to move toward plant roots in response to tension gradients must be of considerable importance, particularly for perennial plants with large developed root systems, because a small distance of movement over a considerable combined length of root system would account for an appreciable volume of water. However, for young plants with a newly developing root system, this movement is so slow that sufficient water for normal growth would not be supplied unless the plant roots are able to extend themselves outward into a fresh soil mois- ture supply. Therefore, as has been described by Davis (^5), when a new corn plant is developing, the available moisture is extracted in the vicinity of the base of the plant and the soil approaches the wilting condition whereas just a few inches further away from the plant the soil may be at or near field capacity. The roots of the newly developing plant must extend themselves outward in order to maintain a continu- ous supply of available water. When the roots of the plant have per- meated the soil region in which they can grow well, the soil moisture tension throughout the soil region occupied by roots will approach that corresponding to the wilting percentage, and, unless additional mois- ture is supplied by rain or irrigation, vegetative growth will cease and the plant will show symptoms of wilting. The depletion of soil mois- ture under perennial plants with developed root systems has been described in detail by Veihmeyer and Hendrickson (12]). During this process of moisture extraction by roots the soil water is withdrawn successively from the large, medium, and fine pores and is replaced by a gaseous phase. Thus, aeration processes are expedited and become freer as the moisture content of the soil is depleted. Soils having favorable structure for the growth of plants tend to have an appreciable fraction of the pore system made up of large pores. These pores are the first to empty during moisture depletion, thus promoting good drainage and favorable aeration. From saturation to the field capacity, moisture depletion is accounted for by downward drainage. Leaching and loss of soluble material from the profile occurs during this process. After the field-capacity condition 416 Mineral Nutrition of Plants is attained, however, moisture depletion is largely brought about by plant root extraction. In this case the soluble materials in the soil solution are either absorbed by the plant during moisture uptake or, as is the case in saline soils, the soluble salts are largely left in the soil because of the selective absorption of the plant root. Moisture depletion is thus accompanied by a considerable change in the concentration of the soil solution. This change in the concentration of the soil solution brings about concurrent changes in the equilibrium between the adsorbed and the soluble ion phases. It was indicated earlier that the freezing point depression gives a measure of the total soil moisture stress. The aqueous vapor pressure of the soil water is another measure of the free energy of the soil water or the total soil moisture stress. In the plant growth moisture range, the pressure of the water vapor in the soil atmosphere is always very high and near the saturation pressure at the particular temperature. The relative humidity range from 98 to 100 per cent more than covers the available moisture range in all soils. In other words, the relative humidity in soil from which plant roots are extracting water for growth never goes below about 98 per cent. MOVEMENT OF MOISTURE IN SOIL The movement of water into and through soil can be conveniently expressed in terms of the force which tends to produce the motion of the water. Gravity and the gradient of the moisture tension in the soil water are the two components that must be considered. Gravity always acts in the downward direction. The component of force arising from the tension gradient in the soil water, however, may act in any direction. When water is at rest under gravity and hence is at static equilibrium, the pressure gradient force is equal to and opposite to gravity. The pres- sure gradient force in soil moisture can be conveniently estimated from tensiometer readings and these instruments cover the tension range in which ready movement of soil moisture takes place. The vector sum of the gravity force and the pressure gradient force can be most conveni- ently expressed in terms of the hydraulic gradient. When the pressure force vanishes, i.e., when the soil moisture tension of the soil water is everywhere the same, then the net water moving force is simply and Wadleigh and Richards 417 entirely gravity force, i.e., "one g," and corresponds to unit hydraulic gradient. Permeability tests on the rate of movement of water in soil cores or soil samples in the laboratory are often conducted at unit hydraulic gradient. When a thin covering layer of water is maintained on the top of the soil column and the water outflow at the bottom of the soil column is at atmospheric pressure, there is no pressure gradient in the soil column and gravity is the only force acting. The velocity of water in soil at unit hydraulic gradient is designated as permeability and is a measure of the readiness with which a soil transmits water. Darcy's law for the movement of water in soils simply states that the velocity of water movement (v) is proportional to the hydraulic gradient (/) as expressed by the equation v = Pi. The proportionality constant, P, is the permeability. The hydraulic head at any point in a soil moisture system is equal to the elevation at which water will stand in a riser or piezometer tube connected to the point in question (101). If the soil is not saturated, a porous cup can be used for establishing connection between water in a manometer and water in the soil, as shown in Figure 2. In practice, mercury manometers are used so as to allow measurements to be made above ground, but actually, water manometers such as those illustrated at the right in the figure could be used. When the porous cup is in unsaturated soil, the free surface of the water in the manometer will come to rest at a level below the porous cup. The difference in level between the cup and the free water surface is a direct measure of the soil moisture tension at the porous cup, whereas the elevation of the free surface referred to any convenient datum is the hydraulic head of the soil water at the cup. The average hydraulic gradient between two points on a flow line, i.e., the soil moisture tension gradient plus the component of gravity force, is equal to the difference in hydraulic head between the points divided by the distance along the flow line. These relations can be illustrated by the tensiometer system shown in Figure 2. Let it be assumed that the manometer readings represent moisture conditions a week after a rain on a fallow soil and, for simplicity, let it further be assumed that moisture conditions do not change in the horizontal direction. The moisture tension at the depth of a porous cup, when expressed in terms of a length of water column, is equal to 418 Mineral Nutrition of Plants the vertical distance from the cup down to the free surface of the water in the manometer. This tension can be read directly from a scale on the mercury manometer {102). Tension readings measure the tenac- ity with which water is held by soil, but are not simply related to mois- n r Lab Lbc -cd fR\ 1 I I V\ \\W\\\V Hab T H cd Figure 2. Porous cup-manometer system for measuring the hydraulic head and hydraulic gradient of water in unsaturated soil. See page 41J. ture movement because gravity force is not taken into account. Hy- draulic head and hydraulic gradient on the other hand do take gravity into account. As defined above, the hydraulic head of the water in the soil adjacent to a porous cup is equal to the elevation of the free water surface in the corresponding water manometer. In a connected soil moisture system, Wadleigh and Richards 419 moisture always streams through soil in the direction of the decrease in hydraulic head. It is apparent that the hydraulic head at cup b is higher than at cup a. Therefore, water is moving upward in the soil layer a— b as indicated by the arrow. In a soil region where the hydraulic head is everywhere the same, water will be at rest under gravity. The tension gradient will be of such size and direction that it will just cancel gravity. The hydraulic head at cups b and c is the same, so that on the average the hydraulic gradient in this interval is zero. The hydraulic head at cup c is higher than at cup d, so that in the interval c— d water is moving downward, as indi- cated by the arrow. The hydraulic gradient is the change in head per unit distance along a flow line. So in the interval a— b we have for the hydraulic gradient /.lb — H.{h/L.th. In the soil interval b— c the average hydraulic gradient is zero, because Hbc is zero. This should not be interpreted to mean that the transmission velocity at b and at c is zero, but rather that there is a soil layer between a and b above which the moisture is moving up and below which the moisture is moving down. In the lower soil interval, ic& = Hcd/Lc6 and is in the opposite direction to /.,,,. Thus, the direction in which water is streaming through soil can be predicted from hydraulic head measurements. Darcy's law for water movement in soil has long been recognized to hold for saturated soils. Its application to unsaturated soils is not so familiar even though the definition of hydraulic head and hydraulic gradient for the unsaturated case as just given are substantially identical to the saturated case. The only difference lies in the fact that in unsatu- rated soils some of the pores are filled with gas and are not available for transmitting water. As the soil dries out and the pore spaces become filled with air, moisture movement must take place in the water films over the sur- face of the soil particles. Thus, the permeability changes with the moisture content of the soil. For example, it is found that for a sandy soil which is saturated with water, the soil transmits 8 cc. of water per square centimeter per hour at unit hydraulic gradient. The permeabil- ity is thus 8 cm. per hour. At a soil moisture tension of 100 cm. of water, which for this soil corresponds approximately to the condition of field 420 Mineral Nutrition of Plants capacity, the permeability is about 0.2 cm. per hour or 1/40 of the value at saturation. At a tension of 200 cm. or approximately 0.2 atmosphere, the permeability drops to 0.0013 cm- Per h°ur which is 1/6000 of the value at saturation. Measurements of this type obtained a number of 015 o X rr Ld Q_ o I >- 010 00 < Ld a. 005 < LT Z> t- < CO 0 o © DATA «SAND ooCLAY d VIRGIN PEAT A v " " B o CULTIVATED A SOURCE LA RICHARDS (1931) it SJ RICHARDS 8 BD WILSON (1938) o CD M ,o □ -Xtt ^^_ o 0 100 200 300 SOIL MOISTURE TENSION - CM. OF WATER Figure 3. Relation of unsaturated permeability of soil to soil moisture tension. The data points illustrate the rapid decrease in the permeability with a small increase in tension. years ago by L. A. Richards for mineral soils (100) and by B. D. Wilson and S. J. Richards (/J5) for peat soils are plotted to a linear scale in Figure 3. At saturation, the permeability of many normal soils will lie in the range from 1-10 cm. per hour. This range is far above the values shown for unsaturated soils, even at the low tensions of 10-20 cm. of water. It is evident that as air replaces water in the soil pores and the Wadleigh and Richards 421 soil moisture tension increases, there is a very rapid decrease in the readiness with which the soil transmits water. Two or three days after a heavy rain or irrigation, tensiometers in field soils usually read in the range from 50-150 cm. of water. It thus appears that field capacity corresponds to the moisture content at which the moisture films are so thin that the unsaturated permeability becomes very small and, com- pared with plant root extraction, further downward drainage can be neglected. It is of interest to consider some soil moisture systems that are related to plant nutrition and experimental plant growth setups. Take for example the small pots in which plants are frequently grown. When the soil in a pot is allowed to come to equilibrium with a water table at the bottom of the pot, we have the condition that the soil moisture tension is zero at the water table. When equilibrium is established the soil moisture tension at each level in the pot is just equal to the vertical distance to the water table. Near the water table the tension is low and the moisture content of the soil is correspondingly high. All the soil pores, except the very largest, will thus be filled with water. As the distance from the water table increases, the larger pores are increasingly filled with air, the soil moisture tension is higher and the moisture films on the soil particles are thinner. This soil, pore-space, moisture system has been nicely pictured by Gardner and Chatelain (47). For a soil water system that is at equilibrium with the water table and which is protected from evaporation, the hydraulic head of the water in the system is uniform throughout; the hydraulic gradient and hence the water-moving force in the system are balanced because the downward force of gravity is just counterbalanced by the increase in tension with height. If evaporation is allowed to take place from the soil surface, this increases the tension gradient over and above that required to overcome gravity and an upward flow takes place. When plants are grown in pots, it is interesting to note that the soil moisture conditions at the bottom of the pot are very similar to those just described. No outflow of water will take place until the soil moisture tension at the bottom of the pot becomes zero, thus downward drainage of water from a pot takes place just as if there were a water table at the bottom of the pot. Consequently, the soil moisture that exists in a shallow pot 422 Mineral Nutrition of Plants when an excess of moisture is applied and drainage is allowed to take place is at a considerably higher level than that which would exist in this same soil in the field. This higher moisture content may consider- ably alter the aeration and the physiological status of the plants, especi- ally in fine-textured soils. When the moisture content of the soil is at field capacity or above, small changes in the hydraulic head can bring about corresponding movement in the water in the soil. However, when the moisture con- tent of the soil is below field capacity, which corresponds to soil mois- ture tensions of the order of 50-150 cm. of water, then the unsaturated permeability of the soil is small, and moisture movement, i.e., the velocity of movement, is restricted even for large hydraulic gradients. When the moisture in soil has become reduced to near the wilting percentage by plant root extraction, it is found that the process of rewetting of the soil follows a fairly regular pattern. In order to get water to move into and through the soil, enough water must be added to the wetted zone to raise the moisture content of that zone above the field capacity. It must be accepted as an experimental fact that there is a comparatively steep moisture gradient in the wetting front. In order for wetting to proceed, the moisture content of the wetting front must be raised to field capacity or above, otherwise the permeability of the soil will be too low to transmit moisture at a rate that will allow the wetting process to proceed at an appreciable rate. When a limited amount of water is added to dry soil, the moisture content of the wetted portion is raised to field capacity and the rest of the soil will remain dry. This phenomenon has been discussed by Shaw (/08), Veihmeyer (120), Colman and Bodman (■?■?), Kirkham and Feng (68), and others. As a consequence of this wetting phenomenon, it appears that for practical purposes the only method for controlling soil moisture during plant growth is to limit the degree of dehydration before the whole root zone is brought to field capacity, or above, by irrigation. Nevertheless, papers continue to appear in the literature reporting experiments in which plants were grown at constant, controlled soil moisture values. The average moisture content for the whole soil volume may have been maintained nearly constant but there is con- siderable doubt in such experiments as to whether the moisture was evenly distributed after limited water applications. Wadleigh and Richards 423 INFLUENCE OF EXCESS MOISTURE ON NUTRIENT SUPPLY AND AVAILABILITY The influence of an excess of water upon nutrient supply and availa- bility in soils may be segregated into three categories, viz: (a) surface erosion, {b) leaching, and (c) presence of a high water table or other conditions maintaining soil in a relatively wet state. Virtually no soil is devoid of the occasional presence of one of these effects, and fre- quently two or all three of them may be operative. Surface erosion Agriculturists in this country have become increasingly aware of the vast losses in soil fertility incurred by surface runoff. This has been so well publicized that further mention seems superfluous. Yet the enormity of this loss mitigates any danger of overemphasis. Fippin (42) presents data to the effect that the Mississippi River carries 475 million tons of silt into the Gulf of Mexico during the average year, and that this silt load contains 4.5 million tons of the exchangeable bases — calcium oxide, magnesium oxide, and potassium oxide — and 1.5 million tons of phosphoric anhydride and nitrogen. This is indeed an enor- mous quantity of plant nutrients. Fippin also calculated that for the year 1939, the average loss of plant nutrients per acre of row crops in the Tennessee River system was "84.6 pounds of calcium, 97.9 pounds of magnesium, 212.2 pounds of potassium, 13.0 pounds of phosphorus, all expressed as oxides, and 23.8 pounds of nitrogen." These figures amply indicate that surface runoff may bring about an alarming drain on plant nutrients present in the relatively more fertile top soil. How- ever, these values are much higher than those reported by Bryant and Slater (23) for two New York soils subjected to seven different kinds of cover. The small losses of nutrients which they found in surface runoff would probably be compensated by contributions from soil formation processes. Leaching Soil management practices inducing infiltration and curbing nu- .trient loss from surface runoff do not necessarily prevent nutrient loss by percolation. The objection might be raised that prevention of surface runoff merely alters the manner in which plant nutrients are removed 424 Mineral Nutrition of Plants from the soil but not the amount. It is certain that liquid water may not move through a soil without carrying solutes with it. A vast amount of data on leaching has been provided by lysimeter studies. Kohnke et al. (7/) have provided an excellent bibliography and summary of the information available. Since soils have virtually no adsorptive capacity for nitrate and chloride ions, these two anions are readily leachable and the quantities of these ions lost by leaching are rather closely related to the quantities present in the soil and the amount of percolate. The sulfate and bicarbonate ions are also readily leachable, whereas phosphate is usually present in leachate in very small quanti- ties, if at all. Loss of nutrient cations by leaching will be conditioned by base ex- change phenomena. Of the exchangeable bases usually found in soils, sodium has a relatively low energy of retention by adsorption surfaces and is most readily leachable. Consequently, it has been readily leached from humid soils, and therefore it is not prevalent in percolate from soils of humid regions in marked contrast to that from soils in arid regions. Potassium also has a relatively low energy of retention by soil colloids, but it may readily become fixed within the crystal lattice of soil clays. Consequently, it usually occurs in the leachate in rather minor quantities. Thus, calcium and magnesium are almost invariably the predominant cations in the leachate from lysimeters. As Kohnke et al. (7/) point out, the actual quantity and quality of nutrient loss by leaching depends on many different factors. Thus, coarse-textured soils permit a greater proportional loss of nutrients than fine-textured soils, and a porous-crumb structure favors greater percolation than a single-grain structure. The type of soil cover may markedly affect nu- trient loss by leaching. Thus, Lyon and Bizzell (yg) found that an uncropped Dunkirk silty clay loam lost 28 times as much nitrogen by leaching as compared to the same soil continuously cropped to grass. There is much evidence to the effect that a vegetative cover lowers nu- trient loss from leaching, by effecting a reduction both in the amount of leachate and in the content of nutrients in the leachate. Dreibelbis (38) reported that drainage from elaborately designed monolith lysimeters containing a Keene silt loam showed an average annual loss of nutrients in pounds per acre as follows: calcium, 19.9; Wadleigh and Richards 425 magnesium, 10.3; potassium, 13.4; manganese, 0.3; nitrogen, 2.6; and sulfur, 17.8. Losses from a Muskinghum silt loam averaged lower for calcium, potassium, and sulfur but higher for nitrogen. Even though some data indicate that nutrient loss by leaching may be large, these observations do not warrant such a conclusion. Kohnke (70) has em- phasized that losses of plant nutrients in drainage are appreciably less than those possible from surface runoff. When fertilizer is applied to a soil and conditions conducive to leach- ing prevail, there is usually an increase in solute in the leachate, but the proportions of the component ions of the solute usually deviate from those in the added fertilizer due to base-exchange reactions. This is well illustrated by the observations of Volk and Bell (126) on lysime- ters filled with Lakeland loamy fine sand having a base-exchange ca- pacity of 3.0 m.e. per 100 g. Various sodium, magnesium, and potassium salts were applied to different lysimeters. Though no calcium was ap- plied, it was invariably the predominant cation in the leachate. Even when 16.39 inches of leachate had been collected from the 4-foot col- umns of soil, only about 1 per cent of the potassium applied was found in the leachate, whereas about 65 per cent of applied sodium and about 30 per cent of applied magnesium appeared in the leachate. Chlorides were recovered almost quantitatively; about 50 per cent applied sulfate appeared in the leachate and, in most instances, more than twice as much nitrate was recovered as was applied. There was practically no difference in the proportions or amounts of cations in the leachate whether the salts added to the soil were chlorides or nitrates, but when sulfate salts were applied, there was a marked decrease in the loss of calcium. The foregoing data emphasize that leaching losses of nutrients may be high on a light soil under heavy rainfall. Thus, the standard treat- ment of Volk and Bell (126) corresponded to an application of 31 pounds of nitrogen per acre, and the leachate from this treatment cor- responded to a loss of 74 pounds of nitrogen per acre from fallow soil over a five-month period. It is commonly observed that nitrogen defi- ciency may readily develop on light soils during wet seasons. There is increasing evidence that these same conditions are conducive to the development of magnesium deficiency. Hester et al. {59) noted that 426 Mineral Nutrition of Plants magnesium deficiency symptoms were quite prevalent on sweet pota- toes and tomatoes growing in sandy soils in New Jersey during the relatively wet season of 1946. They state that extreme magnesium defi- ciency is not likely to occur on this soil except under conditions of high rainfall or heavy potash fertilization. Boynton et al. (16, ly) have noted that magnesium deficiency symptoms are more prevalent in the apple orchards of New York during wet years than dry. The preponderance of evidence indicates that percolation of water through a soil tends to effect the depletion of sodium, calcium, and magnesium ions with relatively little removal of potassium. Jamison (65) reports that potassium is readily leached out of certain sandy soils in Florida in which the exchange capacity is largely provided by organic matter because of the inability of organic adsorbents to fix potassium. It is probable that in most instances losses of nutrients by leaching do not exceed contributions from soil decomposition and nitrogen fix- ation over a long period of time.* High water tables and wet soils. It is evident that under certain conditions serious losses of plant nutrients from the soil may arise from either surface runoff or down- ward percolation. In addition, retention of excess soil moisture within the root zone may have a seriously detrimental effect on the nutrition and health of crop plants. Impervious layers in the subsoil that bring about permanent or even temporary water tables affect soil aeration, and, consequently, root growth, activity of microflora, nutrient avail- ability, and nutrient entry. As mentioned previously, soils with low permeability may also provide adverse conditions within the rhizo- sphere following a heavy rain or after an irrigation. As the soil mois- ture tension becomes lower, the soil pores become increasingly filled with moisture and gaseous interchange is inhibited. Observations on the composition of soil air illustrate the result of this relationship. Furr and Aldrich (45) studied the composition of the soil atmosphere at various depths under irrigated date palms. On irrigating, the soil * Since the completion of this manuscript, a pertinent study by Chapman et al. on nitrogen gains and losses in lysimeters has come to the attention of the authors. See H. D. Chapman, G. F. Liebig, and D. S. Rayner, Hilgardia, 19:57 (1949). Wadleigh and Richards 427 moisture tension at the six-inch depth dropped from 700-800 cm. of water to 11-15 cm. and remained that low for 4 or 5 days after irri- gating. During this period of low moisture tension, oxygen content of the soil atmosphere dropped from about 20 per cent to 5 per cent. However, a concomitant change in carbon dioxide content was not noted. As soon as 100 cm. of water tension developed following irri- gation, there was observed a rapid increase in oxygen content of the soil atmosphere. The oxygen content of soil air at the 30-inch depth also markedly reflected the irrigation cycles, whereas that at the 96-inch depth was little affected, oxygen content persisting continuously at 15- 17 per cent. However, the carbon dioxide content of the soil air at the greater depths maintained itself at about 5 per cent. These data illus- trate very well the interrelationship between soil moisture content and composition of the soil air. They are especially striking since they were obtained on a relatively coarse-textured soil— Indio very fine sandy loam. Boynton (/j) found that fluctuations of ground water in relatively heavy soils was usually fairly well correlated with fluctuations in oxy- gen and carbon dioxide percentages of the soil air, particularly in the strata of soil just above the water table. Interestingly enough, some- times a month or more elapsed between the disappearance of ground water and the rise of oxygen level to a maximum. Boynton and Reuther (14) recorded oxygen percentages in the soil air as low as 0.1 in the second foot of a Dunkirk silty clay loam under an apple tree during March and April. On subsidence of the water table during the summer, the oxygen content at this depth rose to 17-19 per cent, whereas the carbon dioxide content remained at about 3-4 per cent throughout the year. In the fourth foot, the carbon dioxide content of the soil air re- mained 8-10 per cent throughout the summer. Excess soil moisture conditions soil aeration which, in turn, affects mineral nutrition. The essentiality of adequate aeration for the development of healthy roots and vigorous plants of most species has been emphasized many times both in soils (no, 7, 49, 86, 39, 26, 5, 76) and in solution cultures (75, 48, 109, 41). Numerous studies by Steward and co-workers (113, 114), Hoagland and Broyer (60), Lundegardh (78), and Robertson and Wilkins (/05) have amply shown that the rate of nutrient entry into absorbing tissue is conditioned by the rate of respiration of the 428 Mineral Nutrition of Plants absorbing cells, which, in turn, is conditioned by the supply of oxygen. Lawton {j2) grew corn plants on cultures of Clarion loam and Clyde silt loam in which the moisture regime was maintained at near satura- tion in some of the cultures and at or below field capacity in others. The inferior corn plants produced by the soils maintained at near sat- uration were found to have a relatively low percentage composition of nitrogen, phosphorus, and potassium, but there was little consistent effect on content of calcium and magnesium. Modified soil atmospheres resulting from excess soil moisture affect the nutrition of plants not only through reduced oxygen supply but also by increased partial pressure of carbon dioxide. Bradfield (18) has dis- cussed the influence of increasing partial pressure of carbon dioxide within a system containing calcium carbonate. In general, there is an associated increase in activity of both hydrogen and calcium ions. It is conceivable that these effects could markedly affect availability of cal- cium and also other nutrient ions under certain soil conditions. In fact, McGeorge and Breazeale (81, 21) conclude that a supply of carbon di- oxide in the soil is the primary consideration in maintaining phosphate availability in calcareous soils as a result of its modulating effect on the high pH of these soils. On the other hand, Parker {93) found that on a noncalcareous soil, Norfolk sandy loam, there was no con- sistent effect on the calcium and phosphorus content of plants whether the soil was fortified with carbon dioxide, the carbon dioxide removed, or the soil left untreated. Chang and Loomis (28) have presented evi- dence that increasing carbon dioxide content of the air-supplied nu- trient cultures may be toxic per se to plants over and above any effect of inadequacy in oxygen supply. Increasing partial carbon dioxide pressure would effect an increase in bicarbonate ion concentration in the aqueous phase. Harley and Lindner ($6) noted that when certain apple and pear orchards in Wenatchee, Washington, were irrigated for a number of years with water high in bicarbonate (200-360 p.p.m.), they showed a marked decline in vigor and an increase in incidence of chlorosis. A marked improvement was noted when irrigation water low in bicarbonate was substituted for the high bicarbonate water. Calcium carbonate concretions developed on the roots of the affected trees when high bicarbonate water was applied to the calcareous soil present in that Wadleigh and Richards 429 area. It is of interest to note that the chlorosis which develops under these conditions is characterized by an abnormally high potassium and low calcium content of the leaves (74). The evidence appears to indicate that relatively low concentrations of carbon dioxide in the soil air of calcareous soils aid nutrient avail- ability, but that the high carbon dioxide partial pressures which may occur under excessive soil moisture may be deleterious to root activity and nutrient entry. Excessive soil moisture may have an indirect effect on the supply of nutrients to the plant. Oskamp and Batjer (92) observed in their studies of soil conditions in relation to fruit growing in New York that the most unfavorable orchard locations are those in which shallow rooting occurs because of a high water table during certain seasons of the year. Muller (86) also noted that claypan soils maintaining a high moisture content of the subsoil restrict root development of guayule and may even asphyxiate those roots which have penetrated the lower strata prior to the prevalence of excess moisture. That is, shallow and restricted root development means that a given plant has a correspondingly smaller volume of soil to draw upon for nutrients. Furthermore, the moisture reservoir available to the plant is also restricted so that the plant may suffer from drought and be unable to utilize nutrients in the fertile top soil, even though moisture is present in soil at depths within the normal scope of root penetration of the species. Thus, par- adoxically, the prevalence of excess moisture in the soil during the early part of a growing season may seriously intensify the adverse effect of drought later in the season. Under the anaerobic conditions which may prevail in wet, poorly drained soils, there tends to be a decrease in the degree of oxidation of both inorganic and organic constituents. Lawton (72) observed a marked increase in the extractable ferrous iron content of Clarion loam and Clyde silt loam when the soils were compacted or maintained at high moisture content together with an associated decrease in extract- able ferric iron. Anaerobic conditions resulting from waterlogging of soil may also effect an increase (tremendous in some soils) in exchange- able divalent manganese {112, 73). Studies by Fujimoto and Sherman (4]) indicate that the effect of a level of soil moisture on manganese is 430 Mineral Nutrition of Plants complicated by a hydration-dehydration equilibrium. The complex hydrated oxide found in cool, moist soil has a much lower activity than dehydrated divalent manganese. The evidence available is ample to lend weight to the suggestion of Hoffer (6/) that a relatively high accumulation of reduced iron and manganese in the soil under anaero- bic conditions may well be toxic per se to plant roots over and above any effect due strictly to inadequate aeration of the root surfaces. There is some evidence to indicate that the reduced forms of certain organic components prevailing under the anaerobiosis of waterlogged soils are specifically toxic to plants (106, ig, 94), and that these sub- stances are readily oxidized and rendered harmless under aerobic soil conditions. The concept that specific toxicity of certain organic sub- stances in soil has an adverse effect in plant nutrition has been scoffed at many times, but it has been adequately demonstrated by recent investigations (10, 8, 50). Thorn and Smith (113) point out that the anaerobic decomposition of organic matter in waterlogged soils fre- quently produces hydrogen sulfide. This compound is very toxic to roots. The presence of accumulations of carbon dioxide in the atmosphere of waterlogged soils exerts a modulating influence over the activities of Fe+++ and Fe++, and the consequent influence upon plant nutri- tion, over and above the effect of the state of oxidation-reduction in the system. Halvorson (55) points out that ferrous carbonate is very insolu- ble as compared with ferrous bicarbonate which is readily soluble. Thus, the activity of Fe++ in a given soil system is conditioned by oxidation-reduction potential, pH, and partial pressure of carbon di- oxide within the limitations of the amount and kind of ferruginous mineral present. Halvorson's (55) analysis of the obtaining equilibria indicates that anaerobic conditions in an alkaline soil in the presence of a relatively high partial pressure of carbon dioxide may actually bring about a reduction in solubility of iron as compared with the aerobic state. On the other hand, increased partial pressure of carbon dioxide under aerobic soil conditions is actually conducive to solubility of ferric hydrate. There is considerable evidence to support Kliman's (6g) conclusion that iron enters plants mostly in the reduced state. As a consequence, Wadleigh and Richards 431 the iron supply to plants on wet soils may be either hindered or ac- centuated depending on the status of other prevailing conditions (pH, carbon dioxide pressure), and the activity of iron in the soil appears to affect the relative rate of entry of other plant nutrients into the roots. In light of the foregoing, it is of interest to consider the nutritional disturbance known as "lime-induced chlorosis." There is a great deal of evidence (25, 80, 54, 99, 1, 85, 34) that this type of chlorosis, which is associated with a disturbance in iron metabolism within the plant, is accentuated by wet weather or heavy irrigation and is ameliorated when the soil dries. Wet, poorly aerated calcareous soils would be conducive to accumulations of the bicarbonate ion, and it should be recalled that Harley and Lindner (56) observed the development of chlorosis on apple trees when irrigated with high bicarbonate water. However, Reuther and Crawford (99) found no relationship between the carbon dioxide content of the soil atmosphere and the degree of chlorosis of grapefruit when the intensity of the symptoms varied with soil moisture content over an irrigation cycle. Obviously, the primary cause of chlorosis of plants growing on wet calcareous soils has not been resolved, but a consideration of Halvorson's (55) theoretical treatment along with manganese chemistry and HCO3- activity might prove fruitful. The status of both iron and manganese in the soil is intimately related to the prevailing biological activity (55, 73). Hence, the effect of a high level of soil moisture upon the prevailing micro- organisms must be taken into account. As a case in point, Jones and Tio (6y) observed that symptoms of frenching on tobacco associated with low iron content of the plant could be eliminated by: (a) adding ferrous sulfate to the soil, (b) maintaining a relatively low soil tem- perature, or (c) by autoclaving the soil. The interrelationship between iron availability to the plant and activity of microflora in the soil is implicit in their findings. It is apparent, a priori, that wet, poorly drained soils are favorable to the development of anaerobes and inhibitive to aerobes. Since the anaerobes are capable of using oxygen that is in chemical combination with soil components to meet the needs of their life processes, their activity effects a reduction in iron, manganese, and other reducible compounds (7^). It is also known that denitrification takes place rapidly 432 Mineral Nutrition of Plants in waterlogged soils (/J/, fl, 134). There is a rapid loss of applied nitrate under these conditions, but only a fraction of it is recoverable as ammonia. Willis and Sturgis (134) observed that large quantities of nitrogen as ammonia are lost from waterlogged soil high in nitrogen and maintained at a high temperature (ioo° F.), or from soils high in organic matter. These results indicated that such a soil will tend to reach an equilibrium at which it will maintain a low soluble-nitrogen content against losses induced by high temperatures and alkaline re- actions. De and Sarkar (37) found that much of the difference in nitrogen between the amount of nitrate applied to a waterlogged soil and that recoverable as ammonia was due to nitrogen assimilated by the increased population of microorganisms. Wallihan (131) confirmed this and pointed out that this condition explains the relatively low loss of nitrogen when waterlogged soils are drained. He found that there was a relatively rapid rate of nitrate production following drainage of such a soil, providing further evidence that the denitrification process actually prevents excessive losses of an important plant nutrient from waterlogged soils. On the other hand, the nitrogen so stored may be withheld from crop plants, rice for example, growing on such a soil. In fact, Willis and Sturgis (134) have emphasized the generally poor response observed to applications of nitrogen on rice. They attribute much of this effect to loss of ammonia on denitrification, but the com- petition for nitrogen by microorganisms is also undoubtedly involved. A further effect of wet soils on microorganisms arises from the fact that wet soils tend to be cold (6). That is, the higher the moisture content of a soil the higher its heat capacity. This means that wet soils warm up more slowly during the spring months. Since relatively low soil temperatures depress microbiological activity, nutrient availability dependent on this activity will be correspondingly depressed. MINERAL NUTRITION UNDER SOIL MOISTURE VARIATIONS BETWEEN FIELD CAPACITY AND THE WILTING PERCENTAGE The mineral nutrition of plants within the available range (see p. 412) of soil moisture is conditioned by (a) the extent to which growth and, consequently, mineral utilization might be limited by water supply, (b) the effect of change in thickness of the moisture Wad lei gh and Richards 433 films on nutrient availability, and (c) the effect of variations in soil moisture tension upon microbiological activity. The availability of soil moisture within the available range It is at once obvious that regardless of how nearly optimal the level of mineral nutrients and other growth factors may be, growth will be limited by the extent to which water supply to the plant is limited. Hence, the question arises: is water ever limiting to growth within the bounds of the available range of soil moisture? Veihmeyer and Hendrickson {122, 121) have carried out extensive investigations on the availability of soil moisture to tree fruits growing on deep alluvial soils in California. Their observations indicated that for all practical purposes, soil moisture between field capacity and nearly down to the wilting percentage is essentially of equal availability to the plant. It is evident, a priori, that more osmotic work is required for the entry of water into a plant when the water is restrained by a force of 15 atmospheres (approximate retentive force at wilting per- centage) as compared to a retentive force of only 0.1 atmosphere; i.e., ease of entry of water into a plant may change markedly over the available range. The apparent contradiction between the two previous statements is partially ameliorated by the fact that the relationship between moisture content and moisture retention is invariably hyper- bolic (Figure 1); that is, most of the available water is removed from the soil before a marked increase in soil moisture tension develops. Furthermore, whether or not modifications in growth response will be observed with variations in depletion of soil moisture within the available range will depend on (a) the nature of the soil, (b) prevail- ing weather conditions, (c) kind of plant being studied, and (d) the criterion of growth being used. Deep alluvial soils that permit deep root penetration must be con- trasted with shallow soils or dense, impervious soils. Thus, Boynton (12) found a definite decrease in growth of apples on a shallow soil in New York, if the soil moisture content of the surface two feet decreased to the wilting percentage. That is, the moisture reservoir for these shallow-rooted apple trees was virtually exhausted under the stated 434 Mineral Nutrition of Plants condition. Roots of trees on alluvial soils in California may penetrate 20 feet in depth, so that depletion of the soil moisture in the surface two feet of soil, under such conditions, would not indicate exhaustion of the total moisture reservoir to the tree. Veihmeyer and Hendrickson (724) point out that on compact or dense soils, plants may show symp- toms of moisture stress even though "available" water is still present in the soil. They attribute this to the poor permeation of roots in such soils. Aldrich et al. (j) also noted that irrigated Anjou pears on a Meyer clay adobe showed a decrease in growth before the moisture was depleted to the wilting percentage. They regarded the poor root permeation observed in this soil as a partial contributor to this result. It may be shown by a study of moisture sorption curves for different soils that when 75 per cent of the available water is removed, the remaining water may be under a tension of only 1 atmosphere in some soils, but 5 atmospheres in others. Consider the five soils represented in Figure 1. The data are replotted in Figure 4 in such manner that the moisture content at the 15-atmosphere value is designated as "O" available water; the moisture content of the soils at 0. 15-atmosphere tension (approximately field capacity) is designated as 100 per cent available water. In other words, the relative scale is simply the moisture present in excess of the 15-atmosphere percentage divided by the total available range as just defined. When moisture retention curves are plotted on this relative basis, it is quite evident that their respective loci vary considerably, depending on the specific nature of the soil. For example, it is shown in Figure 4 that when the moisture tension reached 1 atmosphere in the Indio sandy loam, about 84 per cent of the "availa- ble moisture" had been removed; whereas, at this same tension, only about 50 per cent of the "available water" was removed from the Olympic clay. It is reasonable to assume that variations in plant growth associated with variations in soil moisture content above the wilting percentage are more likely to prevail on the latter type of soils than on the former. And there is considerable evidence (128, 129, 36, 44, 64) that growth of plants may decrease with a concomitant decrease in soil moisture content of the root zone at moisture levels above the wilting percentage. The status of prevailing weather conditions may be a determinant as Wadleigh and Richards 435 to whether or not decreasing soil moisture content within the "availa- ble range" will affect plant growth (u8, jo). If the level of soil mois- ture approaches the wilting percentage while the plant is subjected to a cool humid environment, an associated decrease in growth is much less likely than if the plant is growing in a hot, dry environment. That 0 I 2 3 4 OLYMPIC CLAY ANTIOCH CLAY YOLO CLAY LOAM HANFORD FINE SANDY LOAM INDIO SANDY LOAM 10 20 30 40 50 60 70 80 90 SOIL MOISTURE- PERCENT OF TOTAL AVAILABLE 100 Figure 4. Relative loci of moisture retention curves presented in Figure 1. is, at the lower levels of "available moisture" the supply of soil moisture in the former case could probably maintain turgescence, whereas it would probably be insufficient in the latter instance. The evidence indicates that some species of plants, e.g., the tree fruits, show no response in productiveness regardless of the level of soil mois- ture maintained above the wilting percentage. There are productive 436 Mineral Nutrition of Plants vineyards in the San Joaquin Valley of California that receive neither rain nor irrigation during the entire growing season. On the other hand, potato fields in the same area are irrigated daily, and successful main- tenance of Ladino clover pastures in this valley requires that the soil moisture content be kept at near-field capacity. The lower soil tempera- tures prevailing in highly moist soils may be a factor in the need for much more frequent irrigation of potatoes and Ladino clover. The criterion adopted as the measure of growth response is involved in the evaluation of the effect of degree of soil moisture depletion on growth. Adams, Veihmeyer, and Brown (2) studied the effect of vari- ous irrigation regimes on growth and productiveness of cotton. The plots maintained at the highest level of soil moisture produced the max- imum vegetative growth of the plants, and vegetative growth declined with decreasing soil moisture reserve at time of irrigation. Yet, there were virtually no differences in yield of seed cotton per acre among these various treatments. Thus, the plants made a physiological re- sponse to increased level of soil moisture supply, but did not provide a corresponding economic return. Hendrickson and Veihmeyer ($y) observed the maximum vegetative growth of peach trees on plots that had the most abundant water supply. However, the production of these trees was not superior to those irrigated less frequently, and the keep- ing quality of the fruit from the frequently irrigated trees was quite inferior. Here again, maintenance of relatively moist soil produced the maximum physiological response as regards vegetation, but the treat- ment was actually an economic liability. Guayule has been observed to decline in vegetative growth as soil moisture depletion prior to irriga- tion is intensified, but rubber production was found to be increased (129,64). The tenability of the concept that all soil moisture above the wilting percentage is "equally available" to plants is conditioned, therefore, by the criteria used in evaluating the results, in addition to the degree of prevalence of modulating factors. Consequently, the question as to whether or not the soil moisture content above the wilting percentage will become limiting to full utilization of nutrients available to the plant is correspondingly involved. Unquestionably, conditions fre- quently prevail in which plant response to fertilization is limited by Wad lei gh and Richards 437 soil moisture supply within the available range, but any generalization would be hazardous, considering the present state of our knowledge of the subject. Nutrient accumulation in plants at various levels of soil moisture supply It would be logical to conclude that under conditions of adequate nutrient supply, plants that are limited in growth by a relatively low level of soil moisture would have a higher content of mineral nutrients than plants under comparable fertility but not limited in growth by moisture supply. Miller and Duley (84) studied this relationship on corn plants. The growth of corn at an "optimum" level of soil moisture was greater than that of comparable plants in soil maintained at a "minimum" level of soil moisture. The nitrogen, phosphorus, and potassium content of the plants was appreciably higher under the "min- imum" soil moisture level, but the reverse relationship was observed for calcium. Emmert (40) found that the smaller tomato plants grown with a relatively low soil moisture supply were higher in nitrogen and potassium content and lower in phosphorus content than the larger control plants grown at an optimum level of soil moisture. Many studies have provided evidence somewhat at variance with that cited above. Haddock* found no consistent variation in nitrate content in petioles of sugar beets subjected to wide differences in irrigation regime. Learner et al.f found that alfalfa irrigated whenever the soil moisture tension reached 0.4 atmosphere at the i-foot depth had a slightly higher nitrogen content than comparable alfalfa on which irri- gation was delayed until moisture tension reached 4.0 atmospheres at the 1 -foot depth. The effect of soil moisture content on activity of nodule bacteria may be involved here. Most experimental evidence shows that for a given level of fertility, decreasing soil moisture supply is associated with a definite increase in nitrogen content of the plant tissue, a definite decrease in potassium content, and a variable effect upon content of phosphorus, calcium, and magnesium (117,116,82,66, *J. L. Haddock. Personal communication of an unpublished report. +R. W. Learner, S. R. Olsen, C. R. Domingo, and C. A. Larson. Personal com- munication of an unpublished report. 438 Mineral Nutrition of Plants 52,53). In other words, it is well established that when growth of plants is limited by soil moisture supply, nitrogen tends to accumulate within the plant because the rate of entry is approximately maintained in con- junction with a decreased rate of utilization in growth processes; but the general tendency for potassium content to be relatively low in plants on the drier soils shows that rate of entry of potassium decreases to a greater degree than does rate of utilization in these slower growing plants. Hence, the availability of potassium to plants may be depressed at the lower soil moisture contents, depending on the nature of the soil. It is of interest to note, however, that Wimmer et al. (/j6) studied nu- trient content of two varieties of sugar beets at different soil moisture levels and reported that one variety showed the conventional decrease in potassium content with decreasing moisture supply, whereas the other variety showed the reverse trend under the same conditions. If this observation is verified, it will indeed be a remarkable case of speci- ficity in ionic entry between two varieties of a given crop. Although the phosphate ion may accumulate in plants limited in growth by low soil moisture supply (84, 66) there is also evidence* that plants so affected may have a relatively low content of phosphate (116, 117,40). Thus, the effect of soil moisture on phosphate nutrition is far less consistent than that observed for nitrogen or potassium. This seems to be indicative of the wide variation among soils in their fixing power for phosphorus as conditioned by soil moisture content. Miller and Duley (84) grew corn on a fertile silt loam from an alluvial bottom along the Missouri River and, Janes' (66) bean plants were grown on an Arredondo loamy sand fertilized with 1200 lb. per acre of 4—7—5. These soils were conducive to phosphate accumulation under low mois- ture supply. McMurtrey et al. (82) grew tobacco on a Colli ngton fine sandy loam fertilized with 750 lb. per acre of 4-8-12 in the row, and found no effect of soil moisture supply on phosphate content of tobacco leaves. Had- dock* grew beets on a calcareous soil, Millville silt loam; Thomas et al. (116) studied tomatoes presumably grown on a Hagerstown silty clay loam that was variously fertilized. These two experiments yielded evi- dence that the phosphate content of plants was reduced by diminishing *J. L. Haddock. Personal communication of an unpublished report. Wadleigh a fid Richards 439 the soil moisture supply. It is quite probable that these various soils differ considerably in their fixing power for phosphate and that this variation is related to the observed effects of soil moisture supply on phosphate content of plants. The available evidence consistently shows magnesium to be relatively high in plants growing under restricted soil moisture supply (u6, 82, 66). This is in line with the inverse tendency for magnesium deficiency to develop in plants during periods of heavy rainfall. Since there is a tendency for the entry of calcium and potassium into plants to vary reciprocally, it could be inferred that the characteristically low potassium content of plants with inadequate soil moisture supply would be accompanied by a relatively high content of calcium. McMur- trey et al. (82) and Thomas et al. {116) found this to be the case on fertilized soils, but the latter investigators found the reverse trend on their unfertilized plots. Miller and Duley (84) and Janes (66) found virtually no effect of soil moisture supply on calcium content of their experimental plants. It is evident, therefore, that the status of other constituents in the soil have a modulating effect on calcium availability under varying soil moisture content. As pointed out in the first part of this paper, diminishing soil mois- ture content effects a concentration of the solutes in the soil solution. Fertilized plants may even intensify this solute concentration as a result of increased rate of moisture extraction. Thus, Jordon et a/.* found that corn plants fertilized with nitrogen during a dry year not only rapidly depleted the soil moisture to the wilting percentage in the sur- face foot of soil, but also to the 3-foot depth. On the other hand, readily available moisture continuously prevailed in the 3 feet of soil under the unfertilized control plants. In a well-fertilized soil subjected to a prolonged dry spell, this solute concentration may in itself inhibit water availability, so that growth on the unfertilized soil may be better than that on which fertilizer was applied. Neff and Potter ($7) noted that newly transplanted tung trees were injured by mineral fertilization during a dry year. Carolus and Woltz (27) found that during four dry seasons in eastern Virginia, the more nitrogen fertilizer they added to *H. V. Jordan, K. D. Laird, and D. D. Ferguson. Personal communication of an unpublished report. 440 Mineral Nutrition of Plants a Sassafras sandy loam, the lower the yield of potatoes. Adding increas- ing amounts of superphosphate had a moderately beneficial effect, how- ever. The effect of these two fertilizing materials on potato yields under these conditions was directly related to their effect on the solute con- tent of the soil solution. Correspondingly, Rahn (96) found that during a "dry" year, fertilization with manure produced a much higher yield of melons than did mineral fertilization, but that during a "wet" year there was no difference in effect on yield from these two sources of fertility. The need for taking into account soil moisture supply in adjusting the fertility program is well recognized in the Hawaiian Islands. Night- ingale {go, 8g) emphasized in his studies on the nitrogen, phosphorus, and potassium nutrition of pineapples that the capacity of the plant to utilize efficiently the available supplies of these nutrients was con- ditioned by soil moisture supply and other environmental factors. Clements and Kubota (31, 32) have developed a technic for following the status of moisture, nutrient, and sugar content of sugar cane over the course of its development, and of adjusting the irrigation and fertility program in accordance with the trend in the status of the plants. Dr. Clements discusses this technic in an accompanying paper. The effect of change in thickness of the moisture films on nutrient availability The discussion in the fore part of this paper pointed out that as the thickness of moisture films on the soil particles decreases, the intensity with which the water is retained on the particles by surface force action increases. Buehrer and Rose (24) discuss the physical properties of ad- sorbed water. The initial layer of water adsorbed on clay particles is presumably held by a "pressure" of several thousand atmospheres (pF 6-7). The characteristics of water retained under such high pressure differ from those of "free" water. There is found to be a tremendous drop in the dielectric constant and presumably a decrease in its solvent power because of the greatly decreased polarity. This implies, therefore, that with diminishing thickness of moisture films, there is a corre- sponding decrease in the proportion of water in the film with normal Wadleigh and Richards 441 solvent properties, i.e., "unbound" water. Reitemeier (98) determined the dissolved ions in solutions extracted from six soils at four moisture contents, and found that the concentration of nitrate and chloride in- creased as the moisture content decreased. He explained this effect on the basis of the existence of "unfree" water in the soil or negative ad- sorption of monovalent anions, or a combination of both. This relative concentration of chloride and nitrate ions in the outer layers of thin moisture films may also be involved in the usually observed accumula- tion of nitrogen in plants on dry soils. In general, Reitemeier found the opposite effect for cations and polyvalent anions; that is, as the thickness of the moisture films decreased, there was a relative dimi- nution of the concentration of these ions in the outer layers. It may be coincidental, but plants subjected to low soil moisture supply also tend to have a relatively low content of these ions (52, 55). The diminished amplitude of the cationic swarm about an adsorption surface under a thinning moisture film appears to be conducive for potassium ions present to enter the lattice of the clay crystal and become fixed. Volk (i2j) found that alternate wetting and drying of soils treated with soluble potassium salts caused rapid fixation of potassium in a nonreplaceable form, and that very little fixation of this kind took place when the soils were kept continuously moist. This has been veri- fied many times (97, 133, 4, ///). Acid soils fix relatively little potassium when moist, but drying effectively increases potassium fixation. Calcare- ous soils fix potassium when moist and the extent of fixation increases on drying. Martin et al. (83) suggest that the effect of drying on potas- sium fixation is that of increased concentration of ions at the adsorption interface, and that dehydration, per se, is not involved. It is logical to conclude that the relatively low potassium content of plants subjected to low soil moisture supply is related to the increased intensity of potas- sium fixation under such conditions. Soil moisture depletion is also conducive to the fixation of phosphorus (88, 119). Trumble (119) concludes that this explains the relatively low phosphorus content of plants under inadequate soil moisture sup- ply. Neller and Comar (88) report that the extent of phosphorus fixa- tion on drying is directly related to the clay content of soils. 442 Mineral Nutrition of Plants Effect of variations in soil moisture tension upon microbiological activity The important role of microorganisms in the mineral nutrition of plants is discussed in an accompanying paper by Dr. Norman. It fol- lows, therefore, that any effect that varying degrees of soil moisture tension would have on microbial activity may result in an indirect effect on mineral nutrition. There have been numerous studies (//, 29, 5/, 132) on the relation of soil moisture to soil microorganisms, but as pointed out by Bhaumik and Clark (9), in most of this work the soil moisture levels were expressed as a percentage of the maximum water-holding capacity. This technic often fails to ensure even moisture distribution throughout the soil sample, or to maintain constant moisture content over the experimental period. Bhaumik and Clark (9) adjusted the moisture tension of samples of five different soils at o, 0.001, 0.01, 0.05, 0.5, and 3.2 atmospheres, and collected the carbon dioxide evolved during the course of incubation. In two soils the peak rate of carbon dioxide production was at 0.5 atmos- phere of moisture tension, and in the other three soils at 0.05 atmos- phere of tension. For all soils, the peak rate of carbon dioxide produc- tion was observed at or very near to the moisture tension at the aeration porosity limit, taken by convention as 0.05 atmosphere. Total carbon dioxide production was actually at a maximum in Thurman sand at the highest moisture tension used, and it was relatively very low at the lowest two levels of moisture tension. On the other hand, total carbon dioxide production for the Wabash silty clay was only slightly less than maximum on the saturated soil, but was relatively low at the highest moisture tension. The diverse effect of moisture tension on microbio- logical activity on these two soil types differing widely in texture is in- deed intriguing. This difference may be partially explained by the enormous increase in population of fungi at 3.2 atmospheres tension, as compared with zero tension in the sample of Thurman sand. Novo- grudsky (9/) studied the rate of nitrification in a chestnut soil as a function of moisture content. No nitrification occurred when only hygroscopic water was present, but it was evident when the moisture content was equal to about \x/z times maximum hygroscopicity (pre- sumably slightly above the wilting percentage). Nitrification reached Wadleigh and Richards 443 its greatest intensity when the upper limit of film water equaled twice the maximum molecular water-holding capacity (presumably twice field capacity); these results tend to be in line with those of Bhaumik and Clark (9) for their heavier soils. Waksman (132) has reviewed the earlier work on the influence of soil moisture on microbiological activity. Different organisms vary as to the optimum soil moisture content for their activity. Thus, nitrifica- tion is at its highest near moisture content of field capacity, and exces- sive quantities of water are much more injurious than too low a mois- ture content. It is quite evident that to whatever extent the mineral nutrition of plants is dependent upon the activity of soil microorga- nisms, soil moisture level will have an indirect effect on nutrition through its influence on soil microbes. Another aspect of the relation of soil moisture to microorganism activity is concerned with minor element nutrition. The mineral nutri- tion of crops on sandy soils in Wisconsin is intimately connected with the organic matter content of these soils. It has been reported to the authors that during the drought year of 1946, boron deficiency symp- toms became prevalent throughout the state. It could have been that reduced microbial activity due to drought caused insufficient minerali- zation of the boron present in the organic matter of the soils. MINERAL NUTRITION IN DRY SOILS During protracted periods of drought, the fertile surface soil may dry to less than the wilting percentage. This may have drastic conse- quences to shallow-rooted crops. Even though deep-rooted plants may obtain adequate moisture under these conditions from the deeper hori- zons, the question immediately arises as to the ability of plants to absorb nutrients from the fertile top soil when it is drier than the wilt- ing percentage. For example, boron deficiency symptoms usually be- come predominant during a drought (g = .8839 exert negative influences. Radiant energy is not significant at this point. In Table III a similar analysis of factors is reported dealing with the green weight of the sheaths making up the sample. It should be re- membered that this factor, while of no particular moment in itself, is very closely correlated with the thickness of the stem and hence is asso- ciated with an element of growth. TABLE III Partial Regressions of Certain Factors on the Green Weight of Sheaths 0 = r373) Factor Sheath moisture Age Light Leaf nitrogen Soil moisture Maximum temperature (°F.) Rate of leaf emergence Total sugars Minimum temperature (°F.) Beta + + + + 7342 4612 3616 4558 1021 1184 0808 0501 0382 "/" Value ** ** 16.52 12.50 1 1 .42 10.61** 4.06** 3.86** 1.97 1 .67 1 .26 R= .8268 458 Mineral Nutrition of Plants The most dominant factor here is sheath moisture. Sunlight asserts itself as another dominant factor. Age is also more important than here- tofore. The nitrogen level assumes importance but curiously enough is negative in its influence, as are soil moisture and maximum tempera- ture. The rate of leaf emergence, total sugars, and minimum tempera- tures are not significantly related to the girth of the stalk. TABLE IV Partial Regression of Certain Factors on the Rate of Leaf Emergence (" = J373) Factor Beta "/" Value Minimum temperature (°F.) ASe. Maximum temperature (°F.) Soil moisture Light Leaf nitrogen Sheath moisture + 2716 15 31** — 3658 M 38** + 2507 12 90** + l633 10 20** + !5°3 7 02** + 1967 6 66** — 0107 32 In Table IV is reported the analysis of the factors affecting the rate of leaf emergence. Leaf production provides the plant with its meri- stems and additionally is a measure of the plant's vigor. As might well be expected, age is a dominant factor and is, of course, negative. Both day and night temperatures are dominant. Soil moisture here assumes a very important role. In fact, it appears that after the plants have used up about two-thirds of the available water there is a striking retarda- tion of leaf emergence and elongation as well. Light intensity and the nitrogen level are also important factors in leaf emergence and vigor. Sheath moisture which has been a dominant factor in all the categories of growth is not significant here. Thus far we have been determining the weight of factors affecting the growth and vigor of sugar cane. Since the sheath moisture and nit- rogen levels are two practical keys to the manipulation of the physio- logical status of the plant, a measure of factors affecting these levels is worth while. In Table V the analysis of factors affecting the level of sheath moisture is given. Harry F. Clements 459 Soil moisture, being the source of plant moisture, is of course primary, but there are other factors which are important in determining the moisture level of the tissues. However, in these studies the soil moisture was maintained well above the wilting point throughout the crop ex- cept toward the end when the crop was put on a drying schedule. Under these circumstances the dominant factor affecting the moisture TABLE V Partial Regressions of Certain Factors on Sheath Moisture (" = 1373) Factor Leaf nitrogen Maximum temperature Age Light Relative humidity Soil moisture Wind velocity Beta "t" Value + • 6464 30.97** + 2409 12 15** — 1825 8 81** — 1 72 1 6 72** + 1245 6 46** + 0800 5 20** + 0727 3 54** R= .846 level is the nitrogen level. The strong positive relation between maxi- mum temperatures of the air and the moisture level may be somewhat surprising but shouldn't be so when it is remembered that low soil temperatures which follow air temperatures may result in actual desic- cation of the plants even though soil moisture is adequate. Light is strongly negative in its influence on tissue moisture, as is age. The hu- midity relationship is strongly positive as it should be, and it appears that this factor has not been given enough attention in the past. In this analysis wind velocities are comparatively minor but are shown to be positive. However, in the Islands high winds are associated with stronger light intensities and lower humidities and, thus, the influence of wind may be masked, especially since the area in which this study was carried on is not a high wind area. In Table VI a similar analysis of factors affecting the nitrogen level is shown. 460 Mineral Nutrition of Plants The outstanding factor affecting the nitrogen level is tissue moisture. Age is also dominant. In fact, after a crop is in its second year of growth, it is difficult to affect the nitrogen level by application of fertilizer. Minimum temperatures exert a strongly negative influence while maximum temperatures exert a weaker positive influence. Soil moisture is also negative but its influence is weak. Light does not seem to affect the nitrogen level. TABLE VI Partial Regressions of Certain Factors on Leaf Nitrogen (» = 1373) Factor Beta "/" Value Sheath moisture + -6334 29.49** Age Minimum temperature (°F.) Maximum temperature (°F.) Soil moisture Light -.2932 — . 1262 + .0858 -.0570 + .0126 R= .832 14.28** 7.06** 4.43** 3-55** .65 It is evident from our analysis so far that the elements of weather have an overwhelming influence on the growth of sugar cane. When the physiological status of a crop can be controlled by maintaining the moisture and carbohydrate levels at known levels for each stage of growth, it is apparent that the yields which are obtained should be maximal for a particular location and should be related, not so much to the so-called fertility of the soil but to the energy level of the atmos- phere as composed of radiant and heat energy. To be sure, we can hope to produce the maximum crop only occasionally, since every storm, every unfavorable circumstance of soil, weed control, or general mismanage- ment of a crop will reduce the likelihood of the maximum achievement by amounts in direct proportion to the sum total of mishaps. We ought, however, to be able to predict what can be achieved in any given area and proceed from this to narrowing the differences between the actual and the theoretical yields. Harry F. Clements 461 In order to develop this equation for sugar cane, I have taken the factors of maximum and minimum temperatures, radiant energy, and the two dominant physiological factors, tissue moisture and age, and have calculated the partial regressions for growth units. Vigor of plants is not directly used here, although it will be remembered that among other factors, moisture and age were dominant influences on factors of plant vigor. The results of this analysis are shown in Table VII. The five factors show a well-balanced assumption of influence on growth units. TABLE VII Partial Regressions of Certain Factors on Growth Units (» = r373) Factor Beta "/" Value Light Sheath moisture Age Maximum temperature (°F.) Min mum temperature (°F.) + + 1888 4217 10.74** 21 .53** + + 3552 2035 2320 18.97** 1 1 .50** 14.22** R= .8617 From these data, the following equation evaluating the weight of each factor on growth units was obtained : E = 0.0820XJ -\- 9.3206X0 — 0.0979X3 -f- 2.8025X4 +3.2272X5 — 1 130.9004 where Xr is the average daily radiant energy expressed as gram calories per square centimeter per day; X2 is the sheath moisture expressed as percentage of green weight; X3 is the age expressed in days; X4 is the maximum daily temperature as degrees Fahrenheit; and X5 is the mini- mum temperature expressed in degrees Fahrenheit. E is the estimated average number of growth units produced per day. When this equation is applied to the pertinent data collected in this study and the average daily growth units are multiplied by the actual age of the crop in months, the results are as shown in Table VIII. The fact that the averages of the two columns are so very close is 462 Mineral Nutrition of Plants encouraging. Discrepancies between items in the columns means that more work has to be done toward achieving greater accuracy. However, the results are sufficiently good to justify projecting the equation to several of the plantations with which I am associated to determine what yields ought to be obtained at each. In arriving at these estimates I used the temperature and light measurements as obtained at each place. Age would, of course, be the same for all the estimates. For the moisture values I used first that level TABLE VIII Actual vs. Calculated Growth Units Plots Actual Calculated A 1580 M54 B 1951 1882 C 2149 1938 D 2124 2045 RA 1 194 J358 RA 1123 1251 RB 1841 199° RC i7T7 x739 RD 1689 1695 Average 1708 1706 which is near the ideal, and second, levels which were actually obtained at each place. The conversion of growth units to tons-cane-per-acre was accom- plished by multiplying the average growth unit per day for the crop by the age of the crop, that is, 24.0 months. The resulting value was converted to tons-cane-per-acre by multiplying by the conversion factor 0.065. The results are shown in Table IX. In the first column are given the theoretical yields for the light and temperatures actually experienced at the nine places listed, using the same normal moisture curve for each. In the second column the same local light and temperature records were used but the actual moisture level measured on a good crop at each place was used. In parentheses, after each of these yields is the actual yield obtained on the field from which the moisture levels were Harry F. Clements 463 taken. The agreement of the two sets of values in the second column is evidence of the validity of the equation used. It is quite clear that the great variation in yields which is obtained is traceable to the climate of the area. The differences between the items of the first column and those in the second represent the difference between the yields which are possible and those actually obtained. There are only two cases where agreement TABLE IX Estimated Yields at Several Locations for Two- Year Crops Normal Moisture, Actual Moisture, Location tons cane per acre tons cane per acre Ewa 122 127 (128)* Waialua 120 95 (101) Paia 127 97 (93) Kihei 135 104 (99) Upolu 124 100 (90) Halelua 106 78 (71) Puakea 109 86 (86) Maulili 79 57 (51) Puuokumau 62 66 (50) *Actual yield at each place appears in parentheses. is good. In others, there is a considerable margin between what is pro- duced and what might be expected. At Paia, Kihei, Upolu, and Halelua the departures are quite large. These four places are very windy. Crops on these fields invariably show low moisture levels or whipped leaves. Partial solution of the problem here appears to lie in the development of windbreaks. Although this is helpful, it is probable that planting the rows of cane closer together with a strong-topped variety will help to get the wind off the field. In some of these areas the soil is very compact and hard. In these areas no matter how frequent the irrigation, the moisture level remains low. Obviously, anything which will loosen these soils and improve their texture and permit a better penetration of roots and irrigation water will help raise the moisture level of the plants and thereby increase 464 Mineral Nutrition of Plants yields. In still others, such as Puuokumau, weed control has not been very successful. I have said very little so far about the fertilizer requirements of the sugar cane crop. It should be very apparent to you that the amounts of mineral nutrients required by a crop are determined very largely by the climate in which the crop is growing. It seems only common sense that a crop growing in an area capable of producing 125 tons of cane per acre will require more plant food than one growing in an area capable of producing 50 tons per acre. The actual fertilizer applied will be the difference between that required by the plant in a given energy level and that available to the crop from the soil and irrigation water. It follows as a corollary that since climate is dominant in determining the quantity of growth obtained in the various areas of the islands, successive yields on a given field will also vary according to the weather actually experienced by each crop in the succession, and that, therefore, the fertilization of a given crop can reasonably be expected to vary from the previous crop. Furthermore, another variety of sugar cane can have a different requirement not only because its needs are different but also because its ability to extract the needed materials from the soil is different. With such a variety of circum- stances affecting the welfare of a crop, it seems obvious to me that the well-being of a crop has to be followed while it grows, and that empiri- cal practices at best can be successful only occasionally. summary: the crop log To bring into focus, then, the requirements of a crop as it grows so that it can be fed, and otherwise nurtured and guided to a successful harvest, the crop log (/, 6, 2, 3) was developed and is now in practical and successful use on over fifty thousand acres of sugar cane. Such a crop log is shown in Figure 1. It is a record of the crop's progress from its start until harvest and is made up of certain physical and chemical measurements and observations which serve as a guide to its handling. In the top section is a record of the maximum and minimum tem- peratures and sunlight experienced by the crop as it grows. Below this is the Growth Index section. If elongation measurements are kept, they are recorded here. Accumulated growth is indicated on the top line. The green weight of the sheath sample per stalk is also recorded. o 1— 'cal 3 O a. E 85 600 4 \ c 75 N V ^L \ / \ / % / 00 K O 65 \ / ft cn 00 ■^""40 238 < 1.9 1 4. 2 5 4 6 5 7 6 8 9 IC .0 II 1 12 3 13 4 14 .6 15 i ie .9 18 0 19 .2 2C .3 21 5 22 7 " X 1 ?"? C c to 2 0 o -. EZ 1.5 u o *V* ^_ ■ 'in ° z 10 £ in r- Cvjl _ m g f CO in CM a> ro CO W UJ C\j HI Csl s O ^ cn if CvJ CD 4- > ki > a a > o XI i_ i_ >J c 3 O o a> a 1 3 a> u o ai a> a o <| O Z Q -» u_ S < -3 -> < CO O z Q u. S < £ -^ Q in (0 -<* 8.5 1 1 111 I 1 ~r T~ ~r 11 J \\ 11 H 111 \ 1 K C Q) 3 8.0 75 2" X t/> Rainfall o 2 r.u 1 i 1 1 1 l X 0) 15 -> V y yr \ x T3 / * s / >* l_ O § 10 .5 - 1 / CL X 01 T3 4 X CD c .3 E 3 o .2 225 □ cs 1 _ O CO ID IT a (A O o CO 10 m JD t_ «_ u £■ l: n o a a 0 3 £ u a a z O Q Lu < -; 0. O Q LL < —3 Crop Year 1947 S torted 6/29/45 HARVEST DATA FERTILIZATION, etc. Field RB c tond (6mos) Ex. TCA 120.0 BRIX 21.0 210 N o P20s 73 K„0 Age at 1st. N 0.63ifros. Acres 0.6 r arvested 6/24/47 TSA 1558 Pol 18.7 Cycle 151 Re toon * ige 23.8 TC/TS7.70 Pur. 89.7 Age last N l0.50mos. Vorlety 32-e 560 1 rosh % TCPAM 5.05 Date 1st. Irrig. 7/3/45 Elevation 20' C lead Cane % TSPAM .656 Totol No. Rounds 31 Soil Type Resld ual Total Rainfall 31.58 Weed Control _ Good Towelling 1st. Yr. — None " 2nd.Yr.-V.LigM Figure i. A completed crop log on sugar cane. 466 Mineral Nutrition of Plants The nitrogen index which is the total organic nitrogen content of the green tissue of the leaf blades expressed in the dry-weight basis is plotted in the third section. Also along this curve are indicated the various appli- cations of fertilizers made to the crop. At the top of this section is a record of the accumulating age of the crop. At the bottom are the actual dates on which sample collections were made. Below this is the section dealing with the moisture index. The actual tissue moisture levels are graphed. At the top, the downward pointing arrows indicate dates of irrigation rounds. At the bottom, the vertical bars indicate the rainfall received. The Primary Index, or the total sugar level of the leaf sheaths is plotted next as percentage of the dry weight. The potassium and phosphorus indices are shown next with the normal line dotted across the space at 2.25 and 0.080, respectively. The values are potassium and phosphorus contents expressed on the basis of the sugar-free dry weight. At the bottom of the log are certain miscellaneous data which make the record for the crop complete. The log once completed becomes a permanent record for that crop. It has several uses: it is an excellent source of research material; it, along with others on the same field, serves to point up long-time trends; it is useful in comparing one field with another and often points up needs for differential preparation or cultivation. But its most important use is in guiding the current crop in its growth and culture to maturity. To show its application I shall very briefly describe its use in handling a ratoon crop. As soon as a field of cane is harvested and the cane lines are reshaped, the first irrigation water is applied. Fertilizer applications are then made. Nitrogen is always applied, and in high production areas 75 pounds of the element are put on. Potash is also applied at once if the previous crop log showed need for potash. If it did not, no potash is applied unless and until the potassium index drops below the normal line. Phosphates usually are not applied to ratoons but only to plant crops. During this early period of a ratoon, irrigations are based on a schedule. The tensiometer (9) is being used to excellent advantage Harry F. Clements 467 here, irrigations generally going on with tensiometer readings of 0.30 atmosphere. This schedule is maintained throughout the growing cycle of the crop. When the young crop is between two and three months of age, its plants are large enough to be sampled. Sampling is continued at 35-day intervals throughout the two years of the crop. From these samples, analyses are made for tissue moisture, nitrogen, potassium, phosphorus, and total sugars. These data are plotted on the log. If all our early operations, that is, weed control, irrigations, and fertilizations, have been well timed and properly done, the moisture index is above 85, the nitrogen index at or above 2 per cent, the primary index below 10 per cent, and the potassium and phosphorus indices above the normal line. These levels should be maintained throughout the first six months. A second application of 75 pounds of nitrogen is made at three to four months of age except in a few cases of high residual soil nitrogen. On these, the nitrogen index remains high and the second application is canceled. Occasionally, more nitrogen is applied if the index drops. If all the indices are maintained at these desirable levels, the growth is excellent. Troubles may be encountered in maintaining the tissue moisture level. If the soil was compacted by the harvest, the moisture level of the next crop will drop even though irrigations are normal. In short, anything which interferes with proper growth of roots will result in the lowering of the moisture level and hence in the reduction of growth. Here we have a few things which can be done. Since nitrogen is a strong ally of moisture, we can apply small amounts of nitrogen repeatedly, we can run smaller streams of irrigation water down the furrows for a longer period of time, or we can insert trash dams in the furrows to effect better penetration of water. Actually, at best these are temporary expedients. It is usually best to correct such difficulties when the field is plowed and planted. Accurate layouts of irrigation furrows, incorporation of organic materials, and the selection of a vigorous rooted variety all help to keep the moisture level up on these soils. If the moisture level is low because of severe exposure to winds, the selection of proper varieties is a considerable help. We know that maintaining the moisture level correctly results in the production of heavy tonnage. To be sure, if the moisture level is 468 Mineral Nutrition of Plants • not maintained because of fertilizer deficiency, that would quickly be picked up on the log and applications would be made. If weed control gets out of hand, the moisture level also drops. If the moisture level is correctly maintained through the first eight to ten months of growth along with a low total sugar level, the cane growth is very heavy and is, of course, likely to be succulent. Lodging of the cane in this condition will result in breakage and killing of stalks, hence, we impose a hardening on such fields. Irrigations are discontinued until the moisture index drops to about 77 per cent. During this period, not only are the stalks hardened, but their sugar content builds up. The stools and very likely also the roots become loaded with sugar. That is, we not only consolidate tonnage gains made to date, but we also prepare the roots for another spurt of growth. When the mid-crop hardening is completed, irrigation is resumed on schedule. Usually the nitrogen index has dropped by this time and the remaining fertilizer to be applied is now calculated and applied. If the potassium index has fallen below the line, potash is also applied. Thus, fortified with carbohydrate material and now with moisture and needed fertilizers, the crop is off for its second season of growth. The moisture level rises to between 80 and 83 per cent. Usually following the lodging of stalks, if the second season fertilization and irrigations are properly timed, a heavy flush of suckers is set off adding to the stalk population in the field. After twelve months no further fertilization is practiced even though deficiencies develop, but irrigations are kept on schedule. Up until the crop is 17 to 18 months of age, our chief concern is the piling up of tonnage. Except for the period of mid-crop hardening, we are not concerned with quality. However, beginning with seven months before harvest and continuing until harvest time, our concern is no longer one of producing tonnage but becomes one of producing quality (8). Thus, at seven months before harvest, every field is put on a weekly sampling basis and the data are plotted on a ripening log. We know that for best quality the tissue moisture level should drop gradually to a level of 72 to 73 per cent. We also have learned that this level must be ap- proached gradually. If during the second season of growth the moisture level has been higher— that is, 82 to 83 per cent— the irrigations are Harry F. Clements 469 discontinued, the tensiometers are removed from the field, and the sheath moisture is used as the guide to further irrigations. Such a crop would not be irrigated until the moisture level dropped to approximately 79 per cent. An irrigation would be applied and a new, lower level set up, say 76 per cent. Obviously, if the moisture level of a crop is low throughout the second season, this crop has been ripening already and would call for a much-reduced ripening period. In any event, when the final moisture level of 73 per cent is reached, the crop is harvested. While the moisture level is dropping, the nitrogen level drops and the primary index rises. However, the dominant criterion of ripened cane is the moisture level of 73 per cent. This must be arrived at gradually. Ripening is not effected if the crop is simply dried out. Now, as should be the case, if the crop log enables us to follow the welfare of the crop, it should enable us to focus our attention on the needs of a crop, and therefore should result in increasing yields. In some cases we have saved substantial outlays in fertilizers. In other cases, however, we have used more fertilizer. The important thing, however, is that we have not only increased the tonnage of cane pro- duced but have improved the quality of that cane with a resulting increase in tons of sugar produced per acre per month. Finally, we are learning from the crop log that if we maintain the moisture levels where they should be through proper timing of irriga- tion as well as fertilization and weed control we approach more closely each time the maximum production possible for the atmospheric energy available to us. REFERENCES 1. Clements, H. F., Hawaiian Planters' Record, 44:201 (1940). 2. , Hawaii Agr. Expt. Sta., Biennial Rpt., 34 (1944-46). 3. , Rpt. Hawaiian Sugar Techn., 6th Annual Meeting (1948). 4. , Akamine, E. K., Shigeura, G., and Isobe, M., Hawaii Agr. Expt. Sta., Biennial Rpt., 108 (1944-46). 5. , and Kubota, T., Hawaiian Planters' Record, 46:17 (1942). 6. , and Kubota, T., Hawaiian Planters' Record, 47:257 (1943). 7. , and Moriguchi, S., Hawaiian Planters' Record, 46:163 (1942). 8. , Shigeura, G., and Akamine, E. K., Hawaii Agr. Expt. Sta., Biennial Rpt., 120 (1946-48). 9. Richards, L. A., and Weaver, L. R., /. Agr. Research, 69:215 (1944). !