31.01.2015 Views

HERBOLOGIA - anubih

HERBOLOGIA - anubih

HERBOLOGIA - anubih

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

UDK 63/66 ISSN 1840-0809<br />

<strong>HERBOLOGIA</strong><br />

An International Journal on Weed Research and Control<br />

Vol. 7, No. 1, April 2006


Issued by: The Academy of Sciences and Arts of Bosnia and Herzegovina<br />

and The Weed Science Society of Bosnia and Herzegovina<br />

Editorial Board<br />

Paolo Barberi (Italy)<br />

Vladimir Borona (Ukraine)<br />

Daniela Chodova (Czech Republic)<br />

Mirha ðikić (B&H)<br />

Aniko Farkas (Hungary)<br />

Azra Hadžić (B&H)<br />

Senka Milanova (Bulgaria)<br />

Ševal Muminović (BiH)<br />

Shamsher S. Narwal (India)<br />

Zvonimir Ostojić (Croatia)<br />

Danijela Petrović (B&H)<br />

Marko Skoko (B&H)<br />

Lidija Stefanović (S&M)<br />

Taib Šarić (B&H)<br />

Dubravka Šoljan B&H)<br />

Editor-in-Chief: Prof. Dr. Taib Šarić<br />

Technical Editor: Dr. Mirha ðikić<br />

Address of the Editorial Board and Administration<br />

Herbološko društvo BiH (Poljoprivredni fakultet)<br />

Sarajevo, Zmaja od Bosne 8, Bosna i Hercegovina<br />

Phone: ++387 33 653 033, Fax: ++387 33 667 429<br />

E-mail: tsaric@bih.net.ba<br />

Published four times a year<br />

The price of a copy of the Journal: 15 €<br />

Printed by<br />

Štamparija GARMOND GRAPHIC, Sarajevo


CONTENTS<br />

Page<br />

1. A Jubilee……………………………………………………………...1<br />

2. V. D. Mircov, I. ðalović, Z. Brocić: Results of weed control in<br />

field potatoes…………………………………………………………3<br />

3. Anikó Farkas: Soil management and tillage possibilities in weed<br />

control………………………………………………………………..9<br />

4. B. Konstantinović, Maja Meseldžija, Dragana Šunjka: Resistance<br />

study of Amaranthus retroflexus L. species population to the<br />

herbicide imazethapyr……………………………………………….31<br />

5. Tsvetanka Dimitrova, Senka Milanova: Influence of the adjuvant<br />

Desh on the efficacy and selectivity of imazamox 40 a.i.l -1 (Pulsar 40)<br />

in three perennial legume crops……………………………………..41<br />

6. Z. Pacanoski: Herbicide-resistant crops - advantages and risks…...47<br />

7. T. Šarić, I. ðalović: Production of allergenic pollen by ragweed<br />

(Ambrosia artemisiifolia L.) is increased in CO 2 -enriched<br />

atmospheres…………………………………………………………59<br />

8. G. Malidža, V. Janjić, I. ðalović: Genetically modified<br />

herbicide–tolerant crops – state and perspectives............................67<br />

Instruction to Authors in Herbologia………………………………….94


Herbologia Vol. 7, No.1, 2006.<br />

A Jubilee<br />

50 years of Prof. Taib Saric’s work in science and higher education<br />

Professor Taib Saric, Editor of Herbologia, this spring<br />

celebrates 50 years of his fruitful and devoted work in agricultural<br />

science and university education. His contribution to agricultural<br />

profession, science and education has been immense. Although<br />

retired, as professor emeritus he continues to give his significant<br />

contribution, particularly in guiding M.S. and Ph.D. candidates, and<br />

in editting the International Journal for Weed Research and Control<br />

Herbologia.<br />

By this modest article we wish to thank our respectable and prominent<br />

Professor for the tremendous work he has done during five decades and to wish<br />

him good health, to stay further with us for many years on the benefit of agricultural<br />

science and practice.<br />

Short curriculum vitae<br />

Profesor Taib Saric was born in Capljina (Bosnia and Herzegovina) in<br />

1934. He took his B.S., M.S. and Ph.D. degrees in agronomy from the University of<br />

Sarajevo. He studied weed science and agroecology at post-master′s study at Kansas<br />

State University, Manhattan, Ks. and Bucknell University in Pennsylvania, U.S.A.,<br />

and post-doctoral study in weed science at the Wageningen Agricultural University<br />

in the Netherlands. He joined the Indian Agricultural Research Institute in New<br />

Delhi for two years performing research on subtropical field crops.<br />

From 1962 he has been working with the Faculty of Agriculture of the<br />

University of Sarajevo as an assistant, assistant professor, associate professor, full<br />

professor and professor emeritus in agroecology, soil management, weed science,<br />

and environmental protection, participating in education of 44 generations of<br />

students.<br />

In the last 30 years he was Editor-in-Chief of three scientific journals, the<br />

last of them being Herbologia. He published more than 100 research papers, about<br />

600 professional articles, and 24 books (with updated editions a total of 44 books)<br />

on soil management, agroecology, weed science, and environmental protection. His<br />

book Soil Management (four editions) was the major textbook for students of<br />

agronomy in Yugoslavia. His Weed Atlas (five editions), with nomenclature in nine<br />

languages, has been used in about 30 countries all over the world.<br />

He was one of the founders of the Weed Science Society of Yugoslavia (in<br />

Sarajevo, 1973) and was its president. He was one of the founders of the European<br />

Weed Research Society (EWRS, in Paris, 1975) and a member of its Scientific<br />

Committee and Educational Committee. During 30 years he was National<br />

Representative of Yugoslavia, and latter on of Bosnia and Herzegovina, to the<br />

EWRS. He led the foundation of the Weed Science Society of Bosnia and<br />

Herzegovina.


A Jubilee<br />

In 1977 at the Faculty of Agriculture in Sarajevo he founded the first<br />

postgraduate study in weed science with teachers from Yugoslavia, Switzerland,<br />

Italy, etc.<br />

He presented numerous papers with results of his research at Symposia and<br />

Congresses held in Uppsala, Moscow, Warshaw, Leipzig, Brno, Bratislava,<br />

Keszhtely (Hungary), Poznan, Brighton, Gent, Sydney, etc. He was chairperson at<br />

the EWRS Symposia in Uppsala, Paris and Lisbon. As a guest of the Indian<br />

Government he delivered the invited introductory lecture on the methods of<br />

promotion of field crop production on the 63 rd Indian Science Congress, Section for<br />

Agriculture, in Vishakhapatnam in 1976. After that, he lectured at graduate studies<br />

at Agricultural Universities in New Delhi, Hyderabad and Bangalore.<br />

He was a foreign peer-reviewer of several Ph.D. thesis at foreign<br />

universities and in electing full professors at them.<br />

For his rich scientific opus and prolific profesional publications Prof. Saric<br />

was conferred the State Award „Veselin Maslesa“. He was twice nominated a<br />

laureate of the World Food Prize (known informally as the Nobel Prize for Food<br />

and Agriculture). He is a member of the Academy of Sciences and Arts of Bosnia<br />

and Herzegovina.<br />

Contribution to weed science<br />

In the course of 50 years, Professor Saric greatly contributed to weed<br />

science through his research, his numerous publications (papers and books) and<br />

through educating many generations of undergraduate, postgraduate and doctoral<br />

students. His research was particularly devoted to spreading of new, invasive weed<br />

species, various ways of weed control, herbicide testing and application, and cropweed<br />

allelopathy. In studying the weed flora of Bosnia and Herzegovina he<br />

continued the research started by Komsa and Vaskovic in the first decades of the<br />

20th century. He pioneered the introduction of herbicides in Bosnia. He was one of<br />

the pioneers of weed science in Yugoslavia. He took part in founding and work of<br />

the EWRS. He was among the organizers of the most of national and international<br />

Weed Symposia and Congresses held in Yugoslavia.<br />

His great knowledge and very rich experience Prof. Saric has all the time<br />

readily transfered to us, his students and coworkers, to other agronomists, and to<br />

farmers. He will for long time remain an unparalleled coryphaeus of our weed<br />

science, as well as the science of soil management and agroecology.<br />

Prof. S. Muminovic<br />

2


Herbologia Vol. 7, No.1, 2006.<br />

RESULTS OF WEED CONTROL IN FIELD POTATOES<br />

Vlad Dragoslav Mircov 1 , Ivica ðalović 2 , Zoran Brocić 3<br />

1 University of Agricultural Sciences and<br />

Veterinary Medicine of Banat – Timisoara, Romania<br />

2 Faculty of Agronomy – Cacak, Serbia and Montenegro<br />

3 Faculty of Agriculturae, Belgrade – Zemun, Serbia and Montenegro<br />

Abstract<br />

Weed competition can reduce yield and potato quality, affecting tuber<br />

size, weight, and quantity. In this research paper it was compared the<br />

efficacy of 10 herbicide treatments for controlling prostrate pigweed, kochia,<br />

and Russian thistle in a low organic coarse–textured soil and to determine<br />

their effect on marketable potato yields.<br />

Study results emphasize the need for good weed control for optimum<br />

potato yields. Metribuzin applied alone or in combination with metolachlor,<br />

pendimethalin, or trifluralin plus EPTC gave excellent broadleaf weed<br />

control and the highest marketable potato yields.<br />

Key words: weeds, weed control, field potatoes.<br />

Introduction<br />

Weed competition can reduce yield (VanGessel and Renner, 1990) and<br />

potato quality (VanGessel and Renner, 1990), affecting tuber size, weight,<br />

and quantity (Nelson and Thoreson, 1981; Wall and Friesen, 1990a, Rosales–<br />

Robles et. al., 1999). Weeds interfere with harvest, causing more potatoes to<br />

be left in the field and increasing mechanical injury (VanGessel and Renner,<br />

1990). If a mixed population of annual weeds is allowed to compete with<br />

potatoes all season, each 10% increase in dry weed biomass causes a 12%<br />

decrease in tuber yield. One redroot pigweed (Amaranthus retroflexus L.) or<br />

barnyardgrass [Echinochloa crus–galli (L.) Beauv.] per meter of row reduced<br />

marketable tuber yield 19 to 33% (VanGessel and Renner, 1990;<br />

Baziramakenga and Leroux, 1998).<br />

The critical period for weed removal in potatoes is about 4 to 6 weeks<br />

after planting. Weeds emerging 4 weeks after planting are suppressed by crop<br />

growth (Thakral et. al., 1989). These weeds may not reduce tuber yield<br />

through competition, but can interfere with harvest operations.<br />

Mechanical cultivation does not remove weeds within the row and may<br />

damage potato plants and reduce yields (Nelson and Giles, 1989). Herbicides<br />

can reduce the number of cultivations required and enhance weed control<br />

particularly during the early season before hilling. Many herbicides are


V.D. Mircov et al.<br />

approved for use on potatoes grown on medium– and fine–textured, high–<br />

organic soils. Relatively little information is available regarding the<br />

effectiveness and safety of herbicides for potatoes grown in low–organic<br />

matter, coarse–textured soils.<br />

The objectives of this research were to compare the efficacy of 10<br />

herbicide treatments for controlling prostrate pigweed, kochia, and Russian<br />

thistle in a low organic, coarse–textured soil and to determine their effect on<br />

marketable potato yields.<br />

Materials and methods<br />

Field trials were conducted over a three–year period from 2002 to<br />

2004 at the experimental station University of Agricultural Sciences and<br />

Veterinary Medicine of Banat – Timisoara in Romania. The soil was a<br />

chernozem. Soils were fertilized according to Timisoara recommendations<br />

based on soil tests (N=0.219 mg/100 g soil, P 2 O 5 =21.4 mg/100 g and<br />

K 2 O=31.6 mg/100g soil). Fields were plowed, disked, leveled, and hilled<br />

prior to planting.<br />

A randomized complete block design with three replications was used.<br />

The distance between rows was 75 cm and the distance between the plants in<br />

a row was 33 cm, density being 40.000 plants per hectare. The seed of class<br />

A was used. Potato sowing on April 14, 2002 (cv. Desiree); April 20, 2003<br />

(cv. Adora); and April 19, 2004 (cv. Adora).<br />

Prostrate pigweed, kochia, and Russian thistle were broadcast seeded at<br />

a rate of 1.0 kg/ha –1 each and harrow incorporated prior to planting. The<br />

chemical designations for the proprietary herbicides evaluated were:<br />

Common name<br />

metolachlor<br />

EPTC<br />

fluorochloridone<br />

metribuzin<br />

pendimethalin<br />

trifluralin<br />

Trade name<br />

Dual<br />

Eptam<br />

Racer (proposed)<br />

Sencor<br />

Prowl<br />

Treflan<br />

Preplant incorporated (PPI) treatments were applied April 14, 2002;<br />

April 16, 2003; and April 19, 2004 and immediately incorporated to a depth<br />

of 4 to 5 cm with a tractor–driven rotary tiller. Preemergence (PRE)<br />

treatments were applied April 30, 2002; April 24, 2003; and April 22, 2004<br />

and immediately incorporated with 1–2 cm of sprinkler–applied water.<br />

Visual evaluations of crop injury and weed control were made July 17,<br />

2002; July 21, 2003; and June 23, 2004. Weed control was based on a 0–to–<br />

4


Results of weed control in field potatoes<br />

100% scale, where 0 = no control and 100 = no living weeds. Infestations<br />

were light throughout the experimental area for all three weeds. Weeds were<br />

hoed from the weed–free controls as needed, beginning one month after<br />

planting and continuing through August.<br />

Potatoes were mechanically harvested with a tractor–driven potato<br />

digger on October 2, 2002; September 22, 2003; and September 19, 2004<br />

from 1,5 m of the center two rows of each plot. The harvested potatoes were<br />

graded to separate marketable tubers.<br />

Tubers that were diseased, less than 28/55 mm were discarded.<br />

Previous research indicates that specific gravity is independent of weed<br />

density, so specific gravity was not measured. Values for weed control and<br />

marketable yield were subjected to analysis of variance, and treatment means<br />

were separated by Fisher’s LSD test at the 5% significance level. There was<br />

no significant year–by–treatment interaction, so data were combined for all<br />

three years.<br />

Results and discussion<br />

Fluorochloridone was the only treatment that injured potato plants<br />

during all three years (data not shown). At both rates, potato plants exhibited<br />

chlorosis along leaf veins, and plants were slightly reduced in size. As the<br />

season progressed, injury symptoms diminished, and there appeared to be no<br />

difference in foliar growth among all treatments at harvest.<br />

All herbicide treatments controlled 100% of prostrate pigweed (table<br />

1). Trifluralin, in combination with metolachlor or with EPTC, controlled<br />

less than 95% of kochia (an average of 89% and 88%, respectively). Adding<br />

metribuzin to trifluralin plus EPTC increased control to 100%. Pendimethalin<br />

alone or in combination with EPTC controlled less than 75% of Russian<br />

thistle.<br />

Adding metribuzin to pendimethalin increased Russian thistle control<br />

to 95%. All other treatments controlled 90% or more of Russian thistle. The<br />

unweeded control yielded the least marketable potatoes of all treatments<br />

(table 1) and produced 61% less than the weed–free control. The greatest<br />

potato tuber yields were noted in plots treated with metribuzin alone or in<br />

combination with metolachlor or pendimethalin. Pendimethalin alone and in<br />

combination with EPTC failed to control Russian thistle, and marketable<br />

potato yields were lowest among treated plots. Adding metribuzin to<br />

pendimethalin increased Russian thistle control to 95%, and increased<br />

marketable tuber yields by 54% over pendimethalin alone. Previous research<br />

has indicated that pendimethalin may have a beneficial effect on potato yields<br />

beyond that of weed control, possibly by inducing deeper rooting (Nelson<br />

and Giles, 1989).<br />

5


V.D. Mircov et al.<br />

Table 1. Prostrate pigweed, kochia, and Russian thistle control, and potato<br />

yields, averaged over three years (2002–2004).<br />

Treatments Timing Rate<br />

Weed control a Marketable b<br />

AMABL KCHSC SASKR potato yield<br />

l/ha –1 % t/ha –1<br />

Trifluralin+metolachlor PPI 0.75+1.5 100 89 95 40.6<br />

Trifluralin+EPTC PPI 0.75+3.0 100 88 95 40.9<br />

Trifluralin+EPTC+metribuzin PPI 0.75+3.0+0.25 100 100 100 42.5<br />

Fluorochloridone PRE 0.25 100 100 90 42.0<br />

Fluorochloridone PRE 0.50 100 100 99 38.5<br />

Pendimethalin PRE 1.0 100 99 69 28.9<br />

Pendimethalin+EPTC PRE 1.0+3.0 100 100 70 33.8<br />

Pendimethalin+metribuzin PRE 1.0+0.25 100 100 95 44.5<br />

Metolachlor+ metribuzin PRE 2.0+0.25 100 100 96 43.3<br />

Metribuzin PRE 0.5 100 100 100 45.4<br />

Weed – free control 100 100 100 43.2<br />

Unweeded control 0 0 0 28.7<br />

Lsd (0.05) 1 5 6 30<br />

a AMABL–prostrate pigweed; KCHSC–kochia; SASKR–Russian thistle;<br />

b Tubers 28/55 in diameter;<br />

Combining pendimethalin and metribuzin did not significantly change<br />

marketable tuber yield as compared with metribuzin alone or the weed–free<br />

control in these experiments. Though fluorochloridone at 0.5 l/ha controlled<br />

all weeds in this study, marketable tuber yields from this treatment were<br />

lower than the weed–free control. The early injury appeared to have a<br />

deleterious effect on the crop, at least at the higher rate.<br />

Controlling prostrate pigweed, kochia, and Russian thistle at the<br />

beginning of the season increased marketable potato yields more than 100%<br />

compared with the unweeded control. Yields were greatest where weeds were<br />

controlled with no injury to the crop. All herbicide treatments in this research<br />

with the best broadleaf weed control and no crop injury contained metribuzin.<br />

Conclusions<br />

Study results emphasize the need for good weed control for optimum<br />

potato yields. Metribuzin applied alone or in combination with metolachlor,<br />

pendimethalin, or trifluralin plus EPTC gave excellent broadleaf weed<br />

control and the highest marketable potato yields.<br />

6


Results of weed control in field potatoes<br />

References<br />

NELSON, D. C. AND J. F. GILES. (1989): Weed management in two potato (Solanum<br />

tuberosum) cultivars using tillage and pendimethalin. Weed Sci. 37:228–232.<br />

NELSON, D. C. AND M. C. THORESON. (1981): Competition between potatoes (Solanum<br />

tuberosum) and weeds. Weed Sci. 29:672–677.<br />

THAKRAL, K. K., M. L. PANDITA, S. C. KHURANA, AND G. KALLOO (1989): Effect<br />

of time of weed removal on growth and yield of potato. Weed Res. 29:33–38.<br />

VANGESSEL, M. J. AND K. A. RENNER (1990): Redroot pigweed (Amaranthus<br />

retroflexus) and barnyardgrass (Echinochloa crus–galli) interference in potatoes<br />

(Solanum tuberosum). Weed Sci. 38:338–343.<br />

WALL, D. A. AND G. H. FRIESEN. (1990a): Effect of duration of green foxtail (Setaria<br />

viridis) competition on potato (Solanum tuberosum) yield. Weed Technol. 4:539–542.<br />

ROSALES – ROBLES, E., J. M. CHANDLER, S. A., SENSEMAN, AND E. P. PROSTKO<br />

(1999): Influence of growth stage and herbicide on post – emergence johnsongrass<br />

(Sorghum halepense L.) control. Weed Tech. 13: 525 – 529.<br />

BAZIRAMAKENGA, R., AND G. D. LEROUX (1998): Economic and interference<br />

threshold densities and quackgrass (Elytrigia repens L.) in potato (Solanum<br />

tuberosum L.). Weed Sci., 46: 176–180.<br />

7


Herbologia Vol.7, No. 1, 2006.<br />

SOIL MANAGEMENT AND TILLAGE POSSIBILITIES IN WEED<br />

CONTROL<br />

Anikó Farkas<br />

Szada, H-2111, Szabadság u. 58.<br />

E-mail: letics@citromail.hu<br />

Abstract<br />

On the basis of the research done on interactions between tillage, soil<br />

condition and weediness, conclusions are listed in three main point and 18<br />

sub-points.<br />

a) Importance of favourable soil condition<br />

1. The weather data of the region affirm the growing frequency of dry<br />

years and the tendency of weather extremes. This shows the necessity<br />

of tillage systems that increase the water absorbing and water retaining<br />

capacity of the soil. Tillage methods that improve and maintain soil<br />

condition may gain prominence. The cover of the soil between crops<br />

and the introduction of crops that have a beneficial effect on yield may<br />

become a necessity.<br />

2. The effect of previous years can be shown by exact soil condition<br />

examinations, and on this basis the methods of improvement can be<br />

planned and implemented. The danger of plough-sole and disk-sole is<br />

present on the soil. The damage can be alleviated by cultural and<br />

biological methods.<br />

3. The disk-sole formed after the stubble-clearing of mustard confirmed<br />

the importance of consideration of the soil humidity. Tillage was<br />

enough to break compaction close to the surface (with the exception of<br />

direct drill), which means that smaller damages can be alleviated.<br />

4. The sowing and germination of oil radish may be influenced by the soil<br />

condition changed by the tillage method under the main crop. The soil<br />

loosening effect of the oil radish did not appear in case of direct<br />

drilling, which probably shows the sensitivity of the crop to the soil<br />

condition.<br />

b) Evaluation of soil condition and nutrition level<br />

1. The role of good soil condition and nutrition level in the reduction of<br />

drought damage was proven again in wheat, maize, and spring barley.<br />

2. The 35-45 cm deep loosened soil condition had the best influence on<br />

yield in the biologically favourable crop rotation system, which was<br />

achieved by loosening combined with disking. In dry vegetation period


Aniko Farkas<br />

the beneficial effect of tillage methods saving the soil structure<br />

(cultivator) is also obvious.<br />

3. In winter wheat and maize the yield limiting effect of direct drill is<br />

explained by weediness and the hindrance of water movement in the<br />

soil of the experimental field, susceptible to sedimentation.<br />

4. The yield of maize sown after oil radish was influenced by the<br />

loosening of soil and the nutrition level. While in winter wheat water<br />

retaining played a significant role in the whole growing season, in the<br />

year of maize there were droughts only in the spring. This shows that<br />

the yield of maize was influenced more by the water retaining effect of<br />

the soil condition influenced by tillage methods than the precipitation in<br />

the growing season.<br />

5. The undisturbed soil condition characteristic of direct drill did not<br />

hinder the utilisation of fertilizer in maize, in a year with average<br />

precipitation. It can be stated that in case of good nutrition level the<br />

yield-reducing effect of compacted or sedimented soil can be alleviated.<br />

c)Evaluation of crop rotation order according to weediness<br />

1. The introduction of mustard in the crop rotation is advantageous<br />

because of its weed-limiting, soil-covering and soil-improving effect.<br />

With mulching at an optimal date the weed-promoting effect of the<br />

catch-crop and the unnecessary water loss can be avoided.<br />

2. On the stubble of mustard and winter wheat weeds characteristic of the<br />

area appeared. The higher weed coverage in summer – and thus the<br />

better timing of plant protection and the reduction of the seed-bank –<br />

was aided by the favourable loosening and humidity of the soil.<br />

3. The soil loosening effect of oil radish was proven by penetration values.<br />

It can be used as a protecting crop in dry years if soil humidity loss is<br />

curbed during the sowing. The crop improved the cultivability of the<br />

soil and was proven to be a good green crop because of its good<br />

coverage, rooting and its weed limiting effect.<br />

4. The good weed-limiting effect of the soil is proven, in accordance with<br />

the literature, which was further increased be the greater water loss of<br />

the soil. This makes the re-evaluaton of the role of ploughing necessary.<br />

5. Tillage methods can be ranked according to their weed-promoting or<br />

weed-limiting effect. Among the same conditions direct drill has a<br />

weed-promoting, while regular soil-turning has a weed-limiting effect.<br />

The weed-promoting or limiting effect of tillage methods without soil<br />

turning (loosening, cultivator treatment, disking) is different in each<br />

crop and at each nutrition level. Loosening is good to curb the life<br />

activity of perennial crops.<br />

10


Soil management and tillage possibilities in weed control<br />

6. The high proportion of Echinochloa crus-galli in the total coverage and<br />

its prominent rank in the order proves its adaptation to the different soil<br />

and nutrition conditions.<br />

7. The disking under spring barley modified the morpho-biological<br />

spectrum. The coverage of the perennial Elymus repens increased,<br />

utilizing well the higher level of nutrients.<br />

8. According to the coverage of Ambrosia artemisiifolia ploughing was<br />

put at first place in the order of tillage methods. The high number of<br />

seeds appearing after the turning of the soil showed greater infection.<br />

9. The competitiveness of Ambrosia artemisiifolia shows a tendency. On<br />

soils with low nutrient supply it is more competitive than other weeds<br />

and crops. It reacted with higher coverage to the low nutrition level. In<br />

case of good nutrition level the weed-limiting effect of cultivated crops<br />

is higher, but the development of competing, less dangerous weeds is<br />

also better, against which plant protection is easier.<br />

Introduction<br />

The analysis of traditional tillage systems has a significant place<br />

among crop production research topics. In the past 20 years several essays<br />

were published on the effect of low tillage and no-till systems on cultivation<br />

factors, mainly overboard. The actuality of the analyses is justified by the<br />

different cost demand, the necessity and difficulty of creating harmony<br />

between the inputs and the more exact knowledge of the plant protection<br />

effects of certain systems. In different soil tillage systems tillage influences<br />

physical condition and weediness just like crop rotation order and other<br />

elements of agrotechnique.<br />

One of the unfavourable effects of tillage is soil compaction that presents<br />

production risk on 1.4 million hectares in Hungary. Tillage originated<br />

compaction can be meliorated by mechanical (loosening) and biological<br />

(plants having favourable effect on soil condition) methods.<br />

Weediness is related to soil utilization, tillage and the professionality<br />

of plant protection. The yield loss due to weeds can reach or exceed the 30%<br />

of the total loss. Due to the damage of weeds, the protection against them is<br />

inevitable.<br />

As a consequence of structural and financing problems the cultural<br />

condition of the soils deteriorated and weeds proliferated, many species are<br />

hard to kill. The problem is not new. As a result of the herbicide utilization of<br />

the previous years resistant biotypes gained prevalence and the tolerance<br />

against chemicals increased.<br />

The realization that herbicides have a negative effect on the<br />

environment and food safety influenced the weed control practice positively.<br />

11


Aniko Farkas<br />

One sign is the ambition to keep the weed coverage under the harmful limit.<br />

The result is the reducing of herbicide use to reasonable levels, that is also<br />

eligible according to the EU and international regulations. The application of<br />

weed control measures can be aided by the requirements of agriculturalenvironmental<br />

programmes and subsidies, the weed controlling effect of soil<br />

tillage, the expansion of reduced crop rotation with catch-crops with a<br />

beneficial effect on soil condition and weed control.<br />

The reasons and factors mentioned above make the comparing<br />

analysis of different soil tillage systems useful.<br />

According to this, the objectives of this research were<br />

1. analysis of weediness changing as a result of different cultivation<br />

methods, with emphasis to the various nutrition levels;<br />

2. analysis of the effect of soil condition forming and changing as a<br />

result of various tillage methods on the weediness, with new<br />

analytical methods;<br />

3. determining whether the low tillage can be complied with the<br />

appropriate control of weeds;<br />

4. reaction of some important weed species to the treatments;<br />

5. conclusion on the basis of the results and formulating the<br />

recommendations that can be used in practice.<br />

Material and method<br />

On the Experimental Station of the predecessor of title of the Szent<br />

István University (GATE) several different cultivation treatments were<br />

compared in a soil tillage trial on the basis of their effects on soil condition,<br />

yield and weediness. The significance of the examinations is increased by the<br />

judgement of the effects of tillage and fertilization and the introduction of<br />

catch-crops into the crop order.<br />

The experimental field is in the Gödöllı hills. The soil (liable to<br />

sedimentation, sandy loam) and the precipitation conditions make the yield<br />

safety fluctuant, and the number of crops that can be cultivated economically<br />

is low. The year 2000 and 2002 were drier than the average, the years 1999<br />

and 2001, on the other hand, had more rain.<br />

In the two factorial, strip small plot trial with four replications (a)<br />

signifies the soil tillage methods, (b) the fertilization treatments. Plot sized<br />

are 5x20 m= 100m 2 . The order of crops is: white mustard in 1999, followed<br />

by winter wheat, after the harvest of wheat oil radish, maize in 2001, spring<br />

barley in 2002. The date of weed surveys is shown in Table 1. Table 2.<br />

includes the data of on the cropping practices in the trial.<br />

12


Soil management and tillage possibilities in weed control<br />

Table 1.: Dates of weed surveys (Gödöllı, D trial)<br />

Year Crop<br />

Date of weed survey<br />

1 2 3<br />

1999 white mustard 05. 27.<br />

08. 22.<br />

(stubble)<br />

2000 winter wheat 04. 20. 05. 16. 06. 16.<br />

2000 oil radish 08. 24.<br />

2001 corn 06. 06 08. 20 09. 27<br />

2002 spring barley 04. 16. 05. 27. 07. 08<br />

Treatments employed in the trial and their levels:<br />

Soil tillage:<br />

a 1 : Ploughing (22-25cm): traditional system with several turns<br />

a 2 : Loosening (35-40cm) + disking (16-20cm): soil condition<br />

improving system<br />

a 3 : Tillage system based on heavy cultivator (16-20cm), low-till and<br />

low cost system<br />

a 4 : Direct drill: no-till, soil condition maintaining system<br />

Fertilization in the autumn:<br />

b 1 : 80 kg N + 60 kg P 2 O 5 + 60 kg K 2 O /ha active agent (low dose<br />

according to the soil supply),<br />

b 2 : 160 kg N + 120 kg P 2 O 5 + 120 kg K 2 O /ha active agent<br />

(optimal dose according to the soil supply).<br />

Table 2.: Cropping practices in the trial (Gödöllı 1998-2002)<br />

Term 1999 2000 2000 2001 2002<br />

Crop<br />

Objective<br />

Species<br />

Sowing time<br />

Harvest time<br />

Vegetation period, days<br />

Preceding crop<br />

White mustard<br />

Mulch<br />

mixed<br />

4. 12<br />

6. 29. (stem.)<br />

79<br />

winter wheat<br />

Winter wheat<br />

Grain<br />

Mv MAGVAS<br />

’99. 10. 28.<br />

7. 13.<br />

258<br />

white mustard<br />

Oil radish<br />

Mulch<br />

mixed<br />

8. 4.<br />

frozen (Oct) mulch<br />

89<br />

winter wheat<br />

Maize<br />

Grain<br />

PR36R10<br />

2001.5. 4.<br />

2000. 10. 15.<br />

143<br />

oil radish maize<br />

Tillage method<br />

Date of tillage<br />

Seedbed preparation<br />

Top-dressing<br />

Plant protection (weed control)<br />

Pests, diseases<br />

Number of plants/m 2 a 1 b 1<br />

a 1 b 2<br />

a 2 b 1<br />

a 2 b 2<br />

a 3 b 1<br />

Ploughing<br />

1998. 10. 28.<br />

4. 10.<br />

-<br />

-<br />

165<br />

162<br />

160<br />

156<br />

158<br />

4 treatments<br />

1999. 9. 28-29.<br />

10. 8.<br />

2000. 3. 23.<br />

2000. 4. 24. Segal 65<br />

WG<br />

-<br />

-<br />

460<br />

566<br />

504<br />

575<br />

480<br />

Disking<br />

2000. 7. 17.<br />

8. 4.<br />

-<br />

-<br />

-<br />

102<br />

111<br />

110<br />

120<br />

110<br />

3 treatments<br />

2000 10.31.<br />

2000.5.14<br />

2001. 5. 18.<br />

Post: Titus plus<br />

-<br />

-<br />

64.300/ha<br />

64.600<br />

64.800<br />

64.880<br />

65.200<br />

Spring barley<br />

Grain<br />

Amulet<br />

2002.03. 22.<br />

2002. 07. 03<br />

104<br />

2 treatments<br />

2002. 03. 16-19.<br />

2002. 03. 21<br />

2002.04. 12.<br />

-<br />

-<br />

180<br />

220<br />

210<br />

240<br />

200<br />

13


Aniko Farkas<br />

a 3 b 2<br />

162<br />

570<br />

116<br />

65.300<br />

210<br />

a 4 b 1<br />

a 4 b 2<br />

160<br />

162<br />

310<br />

390<br />

102<br />

112<br />

63.600<br />

63.900<br />

150<br />

190<br />

Plot m 2 5x20 = 100 100 100 100 100 100<br />

Appearance of shoots<br />

1999 11. 20. 2000.08.10-15. 2001.05.12.<br />

Other:<br />

Fertilization residue nutrients fertilization<br />

1999. 04. 16.<br />

Without<br />

fertilization<br />

according<br />

treatment<br />

09. 18.<br />

to<br />

according<br />

treatments<br />

2000. 10.31.<br />

Methods of examination<br />

Method of weed seed content analysis: On the stubble of the white<br />

mustard the upper 10 cm layer of the soil was analysed on the basis of ten<br />

200 cm 3 samples. Weed seeds were isolated with ZnCl 2 sedimentation<br />

method. The size of the seed-bank determining the potential infection was<br />

determined for 1 m 2 .<br />

Method of weed survey: Weed covering examinations were carried<br />

out with the 1 square meter (modified Balázs-Ujvárosi-type) quadrat method.<br />

In white mustard and its stubble coverage percentage was measured at ten<br />

places each time. Measurements were taken three times in winter wheat,<br />

maize and spring barley, and once in oil radish. Evaluation was carried out by<br />

variance analysis.<br />

Method of yield evaluation: The yield data from the plots was<br />

calculated for one hectare. Data was evaluated by variance analysis.<br />

Method of soil condition analysis: To measure soil resistance the<br />

Daróczi-Lelkes-type PENETRONIK penetrometer was used. The resistance<br />

of the upper 40 cm layer of the soil was measured every 5 cm, together with<br />

humidity.<br />

Evaluation of the reaction of Ambrosia artemisiifolia L.: The<br />

different tillage methods were evaluated and placed in order on the basis of<br />

their Ambrosia artemisiifolia controlling effect.<br />

Demonstration of the direct effect of soil resistance on weediness:<br />

The direct effect on soil resistance on weediness was demonstrated by rank<br />

correlation. Rank correlation was carried out according to the steps given by<br />

SVÁB (1981).<br />

Results<br />

1. Result of the seed content analysis<br />

From the samples taken from the stubble of white mustard it was seen<br />

that the area is infected mainly by annual weeds. The composition of the<br />

seed-bank consisted mainly of T 4 -type species. Less seeds belonged to other<br />

annual and perennial species. Late summer annuals contribute to a diverse<br />

seed-bank in the soil (Table 3.). On the basis of the 20 species the soil is not<br />

to<br />

Residue nutrients<br />

14


Soil management and tillage possibilities in weed control<br />

considered rich in weeds. This tendency is apparent also on the field. The<br />

tillage systems, the simplified crop rotation and the employed agrotechnology<br />

(chemical plant protection) all contribute to the decrease of weed diversity.<br />

As for the life form of seed-bank species, the predominance of warm<br />

demanding species is obvious. Weed seed content per 1 m 2 was 38250,<br />

infection is considered low.<br />

Table 3.: Seed-bank of the soil, Gödöllı, 1999<br />

Life form<br />

Number of Number<br />

species of seeds<br />

%<br />

G 1 1 3 0,39<br />

G 3 1 1 0,13<br />

H 3 2 4 0,52<br />

H 4 1 1 0,13<br />

Perennials 5 9 1,18<br />

T 1 2 28 3,66<br />

T 3 1 1 0,13<br />

T 4 12 727 95,04<br />

Annuals 15 756 98,82<br />

Total 20 765 100<br />

Results of the weed surveys<br />

In white mustard total coverage was average in May (9.31%). The<br />

development of the species of the first aspect (7,14 %) were aided by the<br />

precipitation. T 4 life form species that were characteristic of the area had<br />

many germinated plants, but their total coverage is low (1.66 %). Annuals<br />

contributed to the 98% of the total weed coverage.<br />

On the stubble of mustard weed coverage was much higher (71,02<br />

%)compared to the spring results. Germination after harvest was significant<br />

with its 8.1%, and was the fourth in the order. This means that the time of<br />

mulching of the protective crop and the tillage for the following crop has to<br />

be chosen more carefully in case of wet weather than in an average<br />

vegetation period. The development of T 4 species forming the third aspect<br />

(58.71 %) was aided by high precipitation and high temperature.<br />

There were three monocotyledonous weeds present (Digitaria sanguinalis,<br />

Echinochloa crus-galli, Elymus repens). Their total coverage is 9.23 %, from<br />

which perennials presented 1.32 %. The proportion of monocotylednonous<br />

plants in the total coverage increased from the initial 2.15% to 13%, with<br />

Digitaria sanguinalis and Echinochloa crus-galli playing a significant role.<br />

The number of species was 19 and 28 at the two dates.<br />

15


Aniko Farkas<br />

In April in winter wheat the weeds of the T 1 group dominated in<br />

accordance with the date of the survey. The formerly typical cereal weeds<br />

had a low cover percentage. Due to the favourable weather conditions, the<br />

lack of many competitors, and the slower development of the wheat, many<br />

germinated T 4 plants were found. From among monocotyledonous weeds,<br />

Echinochloa crus-galli was significant. The scope of the presence of T 4<br />

weeds is in accordance with the other experiences in the country, and warns<br />

about the tendency and the necessity of creating a proper crop rotation. In<br />

case plant protection was not effective against these species, in next year’s<br />

intertilled crop the weed problem may increase.<br />

Since total coverage consisted in large part of species sensitive to the<br />

applied active agents, the sensitive and early annuals thinned or disappeared<br />

by May.<br />

By June perennials – especially Elymus repens – had the highest<br />

coverage, but differently at the two nutrient levels. Their role in total<br />

coverage and in determining the differences between tillage methods is<br />

similar to the role of annuals in April.<br />

Considering the June data it can be said that if nutrition is<br />

unfavourable, tillage methods without soil-turning may prove to be weedpromoting.<br />

On the other hand, if nutrition is appropriate, the disadvantage of<br />

these tillage methods disappears. The effect of direct drill is antinomic,<br />

according to the literature, because both its weed controlling and weed<br />

promoting effect was seen.<br />

It is favourable that the coverage of perennials is not too high and<br />

their upsurgence in maize is less expected. Within total coverage the<br />

proportion of monocotyledonous plants was rising.<br />

Because of the dry vegetation period wheat did not tiller properly, and<br />

therefore its weed-limiting effect was less. It is probable that this was the<br />

reason why the weed-controlling effect of ploughing could not manifest.<br />

On the basis of the survey taken in oil radish the predominance of E.<br />

crus-galli was obvious compared to annuals and the total coverage, at<br />

optimal nutrition level. In the average of the four treatments annual<br />

monocotyledonous plants contributed to 74.99% of the total coverage. E.<br />

crus-galli has a special significance, because maize was following oil radish.<br />

In maize Echinochloa crus-galli had a high coverage at the time of<br />

the first survey; compared to the others, with the exception of cultivator<br />

treatment on low nutrition level where the coverage of Elymus repens was<br />

close to 4%. Because of the late survey, in the case of the other species more<br />

T 4 weeds appeared.<br />

At the time of the second survey the coverage was tenfold. E. crusgalli<br />

was again the first. A Digitaria sanguinalis, which was missing in June,<br />

climbed to the second place. The coverage of Ambrosia artemisiifolia also<br />

16


Soil management and tillage possibilities in weed control<br />

increased, just like the cover percentage of E. repens and Convolvulus<br />

arvensis.<br />

By September the weed coverage decreased. Many weeds finished<br />

their life activity and the shadowing of the maize was also acceptable. E.<br />

crus-galli still had a high coverage, but shared its place with the formerly<br />

insignificant Solanum nigrum and A. artemisiifolia.<br />

The spring barley was lacking weeds, which was caused probably by<br />

the low precipitation and the coverage of barley. At the time of the late<br />

survey mainly perennials (E. repens) and T 1 species were present, with S.<br />

media as the most characteristic. The adaptability of A. artemisiifolia is<br />

obvious, since young plants were seen even at this date.<br />

By the time of the second survey the field still had low weed<br />

coverage, with the exception of direct drill treatment with higher nutrition<br />

level. T 4 life forms gained prevalence, with E. crus-galli having the highest<br />

coverage. A. artemisiifolia had an almost similar coverage in the cultivator<br />

treatment at optimal nutrition level.<br />

In July after the harvest the coverage was obviously low, but E. crusgalli<br />

still had the highest coverage.<br />

It can be stated that in the weed population that is poor in species T 4<br />

weeds had the highest significance. This is probably the result of the<br />

cultvation practice of the previous years.<br />

As a result of the weed condition after white mustard sown as an<br />

adjusting crop, the effect of the treatments can be evaluated reliably in winter<br />

wheat. In April and June the interaction of the two factors was apparent, so<br />

nutrition influenced the weed limiting or promoting effect of the cultivation<br />

treatments.<br />

At the first date the disking + loosening combination limited the<br />

development of annuals better than cultivator treatment at low nutrition level.<br />

Since total coverage was determined by annuals this was also true for total<br />

coverage. In case of higher nutrition level weed coverage was higher in<br />

treatments without soil turn, but the difference is not significant, which<br />

means nutrition had a balancing effect.<br />

Weed condition in May was influenced by herbicide treatment to such<br />

an extent that the interaction of treatments and the weed-promoting effect<br />

cannot be demonstrated.<br />

According to the results in June the wed limiting effect of ploughing<br />

is just tendential at minimal nutrition level. Nutrition modifies this and in<br />

case of optimal fertilizer level the other treatments curb weeds more than<br />

direct drill.<br />

Tillage preceding the sowing of oil radish and the development of oil<br />

radish balanced the weed condition. The weed limiting effect of ploughing<br />

17


Aniko Farkas<br />

was obvious at both nutrition levels. The tillage accompanying the sowing of<br />

oil radish had an effect on weediness also later.<br />

In maize the interaction of the tillage and fertilizer treatments could<br />

not be demonstrated. It is obvious that the weed limiting effect of ploughing<br />

could be demonstrated in the average of the three dates. The other treatments<br />

had a different effect depending on the level of nutrition but statistically this<br />

effect is not authentic.<br />

In barley the weed limiting effect of ploughing and the weed<br />

promoting effect of direct drill is showing a tendency.<br />

The most significant weed of our days, Ambrosia artemisiifolia L was<br />

evaluated separately. Its coverage was considerable at both nutrition levels.<br />

In winter wheat and maize it had tendentially higher coverage on soils with<br />

proper nutrition supply. On the other hand, depending on the year, it utilized<br />

low level nutrients better than competitive weeds and the crop. This shows<br />

that A. artemisiifolia can be limited by appropriate fertilization and the<br />

cultivation of weed limiting crops. This is in accordance with the demand for<br />

harmony between resources.<br />

Between the different tillage treatments the favourable or<br />

unfavourable effect on the weed cannot be determined. In the order of<br />

treatments the ploughed soil was the most favourable for A. artemisiifolia.<br />

This means that on the weed-infected field the weed-bank of the soil also aids<br />

the proliferation of the weed through annual ploughing.<br />

2. Yield results<br />

Examinations were carried out in biologically favourable crop rotation<br />

system. The crop order of the trial is not typical because of the introduction<br />

of oil radish. Therefore the results can be used also from the viewpoint of<br />

sustainable crop production and integrated technologies.<br />

The yield was the best in case of soil loosened favourably 35-45 cm<br />

deep, independently from the weather of the year. Somewhat lower yield was<br />

harvested from soil tilled with cultivator, and ploughing was the third in<br />

order. In every case the yield was lowest in direct drill treatment. On similar<br />

soils, where direct drill can be favourable because of its soil protecting<br />

function, the lower yield and its effect of weediness call for careful<br />

consideration. In case of professional and continuous chemical protection<br />

weediness can be reduced, and this may have environmental consequences.<br />

The weed limiting effect of loosening is not satisfactory, with the exception<br />

of perennial weeds, but its favourable effect on the soil provides an economic<br />

advantage of ploughing, that has otherwise a better weed limiting effect.<br />

In winter wheat and spring barley the interaction of the two factors is<br />

not apparent, but fertilization had yield increasing effect in all treatments,<br />

18


Soil management and tillage possibilities in weed control<br />

most strongly in loosening + disking treatment. The drough damage lessening<br />

effect of fertilization was demonstrated in accordance with literature.<br />

The undisturbed soil condition characteristic of direct drill did not<br />

limit the favourable utilization of fertilizers in case of maize, in a year with<br />

average precipitation. It can be stated that the yield limiting effect of<br />

compacted or sedimented soil condition can be reduced.<br />

3. Effect of introducing a catch crop between the main crops<br />

The favourable biological effect of crop rotation was increased both<br />

by white mustard and oil radish. Favourable effect was shown in the better<br />

cultivability of the soil. On given soil that is susceptible to sedimentation the<br />

duration of soil loosening is short and the effect can be lengthened by crops<br />

with a loosening effect.<br />

4. Change of soil condition<br />

A compacted layer forms under the layer of annual ploughing, which<br />

can extend also towards the upper layers. The loosening effect of oil radish<br />

improved the soil condition, and this effect was shown also in the deeper<br />

layers. Under the depth of the basic tillage for the next crop (maize) the<br />

penetration values were higher but the thickness of the plough-sole<br />

decreased.<br />

The flaws of the previous years were demonstrated mainly in the<br />

shallow tillage treatments. The 35-45 cm deep loosening of the soil alleviated<br />

this problem, and no soil resistance above 3 MPa (critical value) was<br />

measured. The loosening effect of oil radish was reduced somewhat by the<br />

disking following loosening.<br />

In soil tilled with cultivator a more compacted layer formed under the<br />

layer in question. The loosening effect of oil radish could be shown down to<br />

25 cm depth.<br />

In direct drill plough-sole shows the previous soil-turning tillages. In<br />

the second year the soil is compacted under the sowing, in the upper 10 cm of<br />

the soil. The loosening effect of oil radish was demonstrated only in the<br />

upper 15 cm layer but this effect disappeared in the next year. It is important<br />

that weeds endure compacted soil condition while in case of cultivated crops<br />

the yield is depressed because of the competition and the limiting effect on<br />

rooting.<br />

5. Results of rank correlation<br />

The basic hypothesis of my work was that the soil condition<br />

influences the development of the weeds and plant organisms directly. It was<br />

assumed that in a certain depth soil condition affects weediness. If treatments<br />

are put in order on the basis of MPa values measured in ascertain depth,<br />

19


Aniko Farkas<br />

weediness is related to this order. According to the calculated values of the<br />

rank correlation the connection of the two orders of the treatments can be<br />

demonstrated only in a few cases. This can be due to the fact that treatments<br />

change the looseness of the soil in different depths and extents and in a crop<br />

the different plant populations are presented in different proportions, thus<br />

affecting the order and rank correlation also. On this basis the analysis of<br />

pure stands with different tillage methods is recommended.<br />

Examining the rank correlation in function with the depth certain laws<br />

can be observed. To determine mathemaical relationship a dot diagram was<br />

created, fitting the sixth degree polinom of EXCEL as a trend line. Different<br />

R 2 values were calculated for the different dates and weed groups, but those<br />

show the strong fit of the polinom. The example is shown on the weed and<br />

soil conditions of winter wheat in Table 4.<br />

Table 4.: Rank correlation depending on sample depth, Gödöllı<br />

R 2<br />

Winter wheat 2000 April May June<br />

*<br />

rankcor 1 0,93 0,36 0,78<br />

** Annual<br />

rankcor 2 0,82 0,91 0,98<br />

rankcor 1 0,89 0,91 0,91<br />

Perennial<br />

rankcor 2<br />

0,83 0,91 0,63<br />

rankcor 1 0,75 0,86 0,97<br />

All<br />

rankcor 2<br />

0,88 0,83 0,63<br />

Key:* at minimal nutrition level, ** at optimal nutrition level<br />

It can be seen in Table 4. that the value calculated on the basis of the<br />

coverage of annual weeds in winter wheat in May is low. This is in<br />

accordance with the fact that because of the chemical plant protection<br />

annuals are almost gone from the field. Among “unnatural” conditions this<br />

law does not manifest. In June the value of perennials and total coverage is<br />

average, this is probably due to the fact that the role perennials played in total<br />

coverage is different in each treatment.<br />

The results justify the necessity of other, similar examinations, despite<br />

the difficulties.<br />

6. New scientific results<br />

On the basis of the analyses of tillage, soil condition and weediness the<br />

following new scietific results were determined.<br />

1. On Gödöllı brown soil the relationship between the favourable<br />

loosening of the root zone and the yield was obvious in dry year. The soil<br />

condition created by ploughing ensures average yield.<br />

2. On soil that is prone to sedimentation the yield reducing effect of<br />

direct drill in winter wheat, maize and spring barley was caused by bad soil<br />

condition and the higher coverage of less susceptible weeds.<br />

20


Soil management and tillage possibilities in weed control<br />

3. Tillage methods were put into order according to their weed-limiting<br />

effect. The weed promoting effect of direct drill and the weed limiting effect<br />

of regular soil-turning was proved together with the modifying effect of crop<br />

rotation order and nutrition level, especially in the case of tillage without<br />

soil-turning (loosening, cultivator treatment, disking).<br />

4. On the basis of the decreasing coverage of Ambrosia artemisiifolia L.<br />

ploughing was put to the first rank as the most favourable tillage method for<br />

the development of the species. The great competitiveness of A. artemisiifolia<br />

had a tendency, especially among low nutrition conditions.<br />

5. Analyses carried out during the 4 years of the experiment made rank<br />

correlation possible. It was determined that the method can be further<br />

improved by calculations per species. Since the competition of species is seen<br />

among cultivation conditions, pure stands and more tillage methods can be<br />

taken into account.<br />

6. The dependence of rank correlation on tillage depth was proven by<br />

fitting a polinom. More analyses are necessary to justify or reject the<br />

accuracy of the method.<br />

Summary<br />

The hypothesis was that soil condition affects weeds as plants<br />

directly. The competition for nutrients also influences the predominance of<br />

plant potential. As a result appropriate soil management can avert not only<br />

the further deterioration of our soils but weeds can also be forced back. The<br />

significance of the research is increased by the examination of protecting<br />

intercrops. The introduction of these plants (originally used as green manure)<br />

enables the biological improvement of soils and weed control. Special<br />

attention was given to the adaptation ability of Ambrosia artemisiifolia. New<br />

methods were employed together with weed surveys and soil resistance<br />

measurements. On the basis of rank correlation other relationships could be<br />

discerned.<br />

The growing probability of the more frequent drought years and the<br />

tendency of extreme weather indicate the necessity of tillage systems that<br />

increase the water absorbing capacity of the soil and that retain the water.<br />

Tillage systems that improve and maintain soil condition should get<br />

into the foreground. The coverage of soil between two main crops also<br />

becomes necessary, and the introduction of crops having a beneficial effect<br />

on soil and yield into the crop rotation also.<br />

The effect of the tillage used in the previous years can be<br />

demonstrated and thus the methods for improvement can be planned and<br />

carried out. The damage can be alleviated by tillage and biological methods.<br />

The disk-sole appearing because of the stubble-clearing of white mustard<br />

21


Aniko Farkas<br />

showed the importance of adaptation to the water content of the soil. To<br />

eliminate the compaction of soil near the surface all tillage methods (with the<br />

exception of direct drilling) were sufficient, thus presenting several<br />

possibilities to improve lesser damages. The loosening effect of the roots of<br />

oil radish was missing in case of direct drilling, which shows that the plant is<br />

susceptible of soil condition.<br />

The importance of good soil condition and nutrition in the reduction<br />

of drought losses was demonstrated again. Yield was affected most<br />

favourable by 35-45 cm deep loosening. In dry vegetation period besides<br />

loosening methods that spared the soil structure (cultivator) also had a<br />

favourable effect. On the soil of the experimental area the yield reducing<br />

effect of direct drilling in wheat and maize can be explained by weeds and<br />

the blocking of water transport.<br />

The yield of maize following oil radish went according to the<br />

loosening of soil and its nutrient supply. The yield of maize was influenced<br />

more by the water retaining capacity of the soil than by the precipitation in<br />

the vegetation period. The undisturbed soil condition of direct drilling did not<br />

hinder the utilization of fertilization in case of maize in a year with average<br />

precipitation. It can be stated that in case of good nutrition supply the yield<br />

reducing effect of compacted or sedimented soil can be diminished.<br />

The mustard in the crop rotation had a favourable effect. By mulching<br />

at the appropriate time we can avoid the weed-promoting effect of the<br />

intercrop and the unnecessary water loss. The greater weed coverage in the<br />

summer periods was promoted by the favourable loose condition and the<br />

water content of the soil, and this way plant protection measures could be<br />

timed and the weed seed base decreased.<br />

The soil loosening effect of oil radish was proved by soil resistance<br />

values. It can be employed as protecting crop in dry years if the reduction of<br />

soil water loss is emphasized during the sowing. Oil radish improved the<br />

cultivability and ensured a good green crop effect.<br />

The direct drilling has a weed promoting effect, while ploughing<br />

reduces the weeds. Tillage methods that do not turn the soil (loosening,<br />

cultivator, disk tillage) have different weed promoting or prohibiting effects<br />

in case of the various plants and nutrition levels.<br />

The development of the Ambrosia artemisiifolia is the best among<br />

conditions created by ploughing (the dormant seeds get into the germination<br />

zone). The great competition ability of Ambrosia artemisiifolia has a<br />

tendency for having a greater coverage in case of lower nutrition level. In<br />

case of better nutrition the weed-limiting effect of cultivated crops is higher,<br />

and plant protection is easier to apply.<br />

22


Soil management and tillage possibilities in weed control<br />

The accuracy of the rank correlation method still needs further<br />

analysis. In the future the evaluation of “pure” weed stands and the analysis<br />

of data separated according to species and life forms are expected.<br />

Acknowledgements. The research programs supported by OTKA 32851 and OTKA 34274<br />

References<br />

ANTAL J. (1987): Növénytermesztık zsebkönyve. 2. átdolg., bıvített kiadás, Budapest:<br />

Mezıgazda Kiadó, 507 p.<br />

ANTAL J. (1992): Kettıs termesztés. 833-836. p. In: BOCZ E. (Szerk.): Szántóföldi<br />

növénytermesztés. Budapest: Mezıgazda Kiadó, 887 p.<br />

ANTAL J. (2000): Növénytermesztık zsebkönyve. 3. átdold. Kiadás, Budapest: Mezıgazda<br />

Kiadó, 391. p.<br />

ÁNGYÁN J. és MENYHÉRT Z. (Szerk.) (1997): Alkalmazkodó növénytermesztés, ésszerő<br />

környezetgazdálkodás. Budapest: Mezıgazdasági Szaktudás Kiadó, 414. p.<br />

BAJAI J. (1971): İszi búza talajelıkészítési kísérletek különféle elıvetemények után. 243-<br />

250. p. In: BAJAI J. (Szerk.): Búzatermesztési kísérletek 1960-70. Budapest:<br />

Akadémiai Kiadó, 641. p.<br />

BAJAI J. (1978): A mustár termesztése. 95-100. p. In: ANTAL J. (Szerk.): Olajnövények<br />

termesztése Budapest: Mezıgazdasági Kiadó, 139. p.<br />

BÀRBERI, P. és L. O. CASCIO, B. (2001): Long-term tillage and crop rotation effects on<br />

weed seedbank size and composition. Weed res. 41 (4) 325-340. p.<br />

BASTIAANS, L., PAOLINI, R., BAUMANN, D. T. (2002): Integrated Crop Management:<br />

Opportunities and limitations for prevention of weed problems. 12 th EWRS<br />

Symposium, Arnhem, Wageningen. 8-9. p.<br />

BENCZE J. (1954): Iregszemcse, Pusztapó, Bánkút mezıségi talajainak<br />

gyommagfertızöttsége. Agrártud. Egyet. Kar. Kiadv. I. Bp.<br />

BERECZ K., KISMÁNYOKI T., DEBRECZENI B-NÉ (2003): Szervesanyag<br />

visszapótlással kombinált ásványi N trágyázás hatásának vizsgálata tartamtrágyázási<br />

tenyészedényes modellkísérletekben. In: CSORBA ZS., JOLÁNKAI P.,<br />

SZÖLLİSI G. (Szerk.): III. Növénytermesztési Tud. Nap „Szántóföldi növények<br />

tápanyagellátása” Budapest: Akaprint Kiadó 358-363. p.<br />

BÉRES I. és SÁRDI K. (1994): A mőtrágyák hatása az ıszi búza és két gabona gyomnövény<br />

csírázására. Növénytermelés 43 (3) 211-220. p.<br />

BERZSENYI Z. (2000): A gyomszabályozás módszerei. 334-379. p. In: HUNYADI K.,<br />

BÉRES I,. KAZINCZI G. (Szerk.) (2000): Gyomnövények, gyomirtás,<br />

gyombiológia. Budapest: Mezıgazda Kiadó, 630 p.<br />

BERZSENYI Z. és GYİRFFY B. (1995): Különbözı növénytermesztési tényezık hatása a<br />

kukorica termésére és termésstabilitására. Növénytermelés 44 (5-6) 507-517. p.<br />

BIRKÁS M. (1995): Energiatakarékos, talajvédı és kímélı talajmővelés. GATE, Egyetemi<br />

jegyzet 150. p.<br />

BIRKÁS M. (2002): Környezetkímélı és energiatakarékos talajmővelés Budapest: Akaprint<br />

Kiadó 345 p.<br />

BIRKÁS M. (2003): Talajkímélı mővelés és környezetvédelem. Gyakorlati Agrofórum,<br />

Extra 3.<br />

BIRKÁS M., CSÍK L., STINGLI A., PERCZE A. (2003): A talajmővelés minısítése a<br />

környezeti hatásai alapján. In: CSORBA ZS., JOLÁNKAI P., SZÖLLİSI G.<br />

23


Aniko Farkas<br />

(Szerk.): III. Növénytermesztési Tud. Nap. „A szántóföldi növények<br />

tápanyagellátása.” Budapest: Akaprint Kiadó, 43-48. p.<br />

BIRKÁS M. és GYURICZA CS. (2001): A szélsıséges csapadékellátottság hatása az ıszi<br />

búza néhány termesztési tényezıjére barna erdıtalajon. Növénytermelés 50 (2-3)<br />

333-344. p.<br />

BIRKÁS M., GYURICZA CS., PERCZE A. (2001): A javító, fenntartó és kímélı<br />

talajmővelés. In: JÁVOR A. és SZEMÁN L. (Szerk.): Innováció, a tudomány és a<br />

gyakorlat egysége az ezredforduló agráriumában. Gödöllı, 45-52. p.<br />

BIRKÁS M., NYÁRAI H.F., SZALAI T., SZABÓ M. (1996): A fenntartható<br />

növénytermesztés talajhasználati és talajmővelési alapjai. MTA Szabolcs-Szatmár-<br />

Bereg Megyei Tud. Testülete Tud. Ülése, Nyíregyháza 1996. szept. 28. Kiadvány I.<br />

Agrártud. Szekció p.7.<br />

BIRKÁS M., PERCZE A., GYURICZA CS., SZALAI T. (1998): İszi búza direktvetéses<br />

tartamkísérletek eredményei barna erdıtalajon. Növénytermelés 47 (2) 181-198. p.<br />

BIRKÁS M., SZALAI T., NYÁRAI H. F., FENYVES T., PERCZE A. (1997): Kukorica<br />

direktvetéses tartamkísérletek eredményei barna erdıtalajon. Növénytermelés 46 (4)<br />

413-428. p.<br />

BOGDAN, I., GUS, P., SARPE, N., RUSU, T., HATEGAN, M. (2003): Weeds control<br />

strategies in maize crop from Cluj-Napoca area. EWRS 7 th Mediterranean<br />

Symposium, Adana, 2003. máj. 6-9. 109-110. p.<br />

CLAUPEIN, W. (1994). Möglichkeiten und Grenzen der Extensivierung im Ackerbau.<br />

Triade-Verlag, Göthingen, 224. p.<br />

CSAJBÓK J. (2001): Eltérı befektetési szintő növénytermesztési modellek agronómiai<br />

értékelése. In: JÁVOR A. és SZEMÁN L. (Szerk.): Innováció, a tudomány és a<br />

gyakorlat egysége az ezredforduló agráriumában. Gödöllı, 2001. máj. 17-18. 73-<br />

76. p.<br />

CSAVAJDA É. (2002): Néhány, a talajtermékenység szempontjából értékes<br />

zöldtrágyanövény termıképességének vizsgálata I. A vizsgált növények tömege és<br />

szárazanyag-tartalma. Acta Agronomica Óváriensis, 44 (2) 121-133. p.<br />

CSONTOS P. (2001): A természetes magbank kutatásainak módszerei. Synbiologia<br />

Hungarica 4. Budapest: Scientia Kiadó, 155. p.<br />

CZIMBER GY. (1993a): Északnyugat-Magyarország szegetális gyomvegetációja I. A<br />

Szigetköz búzavetéseinek gyomnövényzete. Növénytermelés 42 (2) 143-154. p.<br />

CZIMBER GY. (1993b): A Szigetköz szegetális gyomvegetációja II. Kukoricavetések<br />

gyomnövényzete. Növénytermelés 42 (3) 241-252. p.<br />

CZIMBER GY. (1993c): Északnyugat-Magyarország szegetális gyomvegetációja III. A<br />

Szigetköz cukorrépa vetéseinek gyomnövényzete. Növénytermelés 42 (5) 409-418.<br />

p.<br />

CZIMBER GY. és HARTMANN F. (1994): A köles nemzetség (Panicum L.) Agrofórum,<br />

5.(5) 26-32. p.<br />

CZIMBER GY., KARAMÁN J., TAMÁS I. (1994): A selyemmályva (Abutilon theophrasti).<br />

Agrofórum, 5 (6) 18-27.<br />

DARÓCZI S. és LELKES J. (1999): A szarvasi PENETRONIK talajvizsgáló nyomószonda<br />

alkalmazása. Gyakorlati Agrofórum 10 (7) 16-18. p.<br />

ECONOMOU, G., AVGOULAS, C., TRAVLOS, I., TRIHA, A. (2003): Comparative life<br />

history traits of Conyza albida populations in Greece. EWRS 7 th Mediterranean<br />

Symposium, Adana, 2003. máj. 6-9. 145-146. p.<br />

FÖLDESI D. (1990): Mustár. 119-123. p. In: HORNOK L. (Szerk..): Gyógynövények<br />

termesztése és feldolgozása. Budapest: Mezıgazdasági Kiadó, 356. p.<br />

GECSE M. (2001): Talajállapot-változások évente ismételt mővelés hatására.<br />

Növénytermelés 50 (1) 83-94.<br />

24


Soil management and tillage possibilities in weed control<br />

GYURICZA CS. (2000): Az értékırzı és hagyományos talajmővelés egyes fizikai és<br />

biológiai hatásainak értékelése. Doktori (PhD) értekezés, Gödöllı, 148. p.<br />

GYURICZA CS., FARKAS CS., BARÁTH CS-né, BIRKÁS M, MURÁNYI A. (1998): A<br />

penetrációs ellenállás vizsgálata talajmővelési tartamkísérletben gödöllıi barna<br />

erdıtalajon. Növénytermelés 47 (2) 199-212. p.<br />

HARANGOZÓ K. (Szerk.) (1988): Az egynyári szálastakarmányok termesztése és<br />

hasznosítása. Budapest: Mezıgazdasági Kiadó, 258 p.<br />

HILTBRUNNER, J., STAMP, P., STREIT, B. (2004): Einfluss verschiedener Leguminosen<br />

auf die Unkrautpopulationen in Winterweizen-Lebendmulchsystemen mit<br />

Direktsaat von Winterweizen im Ökolandbau. Z. PflKrankh. PflSchutz, Sonderheft,<br />

XIX, 517-525. p.<br />

HUNYADI K. (Szerk.): 1988. Szántóföldi gyomnövények és biológiájuk. Budapest:<br />

Mezıgazdasági Kiadó, 484 p.<br />

HUNYADI K., BÉRES I., KAZINCZI G. (Szerk.) (2000): Gyomnövények, gyomirtás,<br />

gyombiológia. Budapest: Mezıgazda Kiadó, 630 p.<br />

IOBC Technikai Irányelvek III. In: MIHÁLY B. és KISS J. (Ford.) (2000): Az integrált<br />

szántóföldi növénytermesztés európai irányelvei. Növényvédelem 36 (különszám),<br />

25-38. p.<br />

JOLÁNKAI M. (2003a): A talajmővelés. Gyakorlati Agrofórum, Extra 3. 2. p.<br />

JOLÁNKAI M. (2003b): Tápanyag-visszapótlás, tápanyagellátás a növénytermesztésben. In:<br />

III. Növénytermesztési Tud. Nap „Szántóföldi növények tápanyagellátása” (Szerk.<br />

CSORBA ZS., JOLÁNKAI P., SZÖLLİSI G.) Budapest: Akaprint Kiadó, 16-21. p<br />

JOLÁNKAI M., SZENTPÉTERY ZS., SZÖLLİSI G. (2003): Az évjárat hatása az ıszi búza<br />

termésére és minıségére. „Agro – 21” Füzetek. (31) 74-82. p.<br />

JÓZSA L. (Szerk.) (1985): Másodvetéső szántóföldi növények termesztése. Budapest:<br />

Mezıgazdasági Kiadó, 256 p.<br />

JÓZSEF CS. és RADVÁNY B. (1988): A kalászosok gyomirtása fokozott figyelmet igényel.<br />

Gyakorlati Agrofórum 9 (4) 54-56. p.<br />

KAHNT, G. (1986): Zöldtrágyázás. Budapest: Mezıgazdasági Kiadó, 139 p.<br />

KAZINCZI G., HUNYADI K., LUKÁCS D. (1998): A contribution to the biology of<br />

cleavers (Galium aparine L.) Z. PflKrankh. PflSchutz, Sonderheft, XVI, 83-90. p.<br />

KÁDÁR A. (2000): A vegyszeres gyomirtás története és várható alakulása. In: KUROLI G.,<br />

BALÁZS K., SZEMESSY Á. (Szerk.): 46. Növényvéd. Tud. Napok. Összefogl.<br />

kiadv. Budapest, 33. p<br />

KÁDÁR I., ELEK É., FEKETE A. (1983): Összefüggés-vizsgálatok néhány talajtulajdonság,<br />

a mőtrágyázás, valamint a termesztett növények jellemzıi között. Agrokémia és<br />

talajtan. 32. (1-2) 57-76. p.<br />

KEMENESY E. (1972): Földmővelés – Talajerıgazdálkodás. Budapest: Akadémiai Kiadó,<br />

427. p.<br />

KISMÁNYOKY T. (1993a): Földmővelési rendszerek. 405-420. p. In: NYÍRI L. (Szerk.):<br />

Földmőveléstan. Budapest: Mezıgazda Kiadó 438. p.<br />

KISMÁNYOKY T. (1993b): Szervestrágyázás. 203-236. p. In: NYÍRI L. (Szerk.):<br />

Földmőveléstan. Budapest: Mezıgazda Kiadó 438. p.<br />

KISMÁNYOKY T. és DEBRECZENI K. (2002): Búza és kukorica mőtrágyázásának<br />

tapasztalatai az országos mőtrágyázási tartamkísérletben. In. PEPÓ P. és<br />

JOLÁNKAI M. (Szerk.): II. Növénytermesztési tudományos Nap. „Integrációs<br />

feladatok a hazai növénytermesztésben” 133-137. p.<br />

KORRES, N. E. és FROUD-WILLIAMS, R. J. (2002): Effects of winter wheat cultivars and<br />

seed rate on the biological characteristics of naturally occuring weed flora. Weed<br />

Res. 42 (6) 417-428. p.<br />

25


Aniko Farkas<br />

KOVÁCS CS. ZS. és NAGY Z. (2003): Komoly kihívás az olajretek termesztése. Gyakorlati<br />

Agrofórum 14 (1) 47-49. p.<br />

KOVÁCS J. (1998): A növényvédıszer-forgalmazás, -értékesítés ökonómiai megközelítése<br />

különös tekintettel a gyomirtásra. Növényvédelem 34 (5) 261-262. p.<br />

KOVÁCS M. (1986): Növényföldrajz és növényökológia. GATE, Egyetemi jegyzet. 185 p.<br />

KOVÁCS P. (1999): Növényvédıszer vásárlás és felhasználás a közelmúltban, különös<br />

tekintettel a gyomirtásra. Gyakorlati Agrofórum 10 (5) 60-61.p.<br />

KÖNNECKE, G. (1969): Vetésforgók. Budapest: Mezıgazdasági Kiadó<br />

LÁSZTITY B., KÁDÁR I., ELEK É. (1981): Mőtrágyázás hatása az ıszi búza<br />

tápelemfelvételére barna erdıtalajon. Agrokémia és talajtan. 30 (1-2) 25-36. p.<br />

LEHOCZKY É. (1994): A gyomnövények és a kultúrnövények versengése a tápanyagokért<br />

355-360. p. In: DEBRECZENI B. és DEBRECZENI B-NÉ (Szerk): Trágyázási<br />

kutatások 1960-1990. Budapest: Akadémiai Kiadó 411 p.<br />

LEHOCZKY É. (2002): Az Echinochloa crus-galli (L.) P. B. és a kukorica korai<br />

kompetíciójának hatása II. A növények tápanyagfelvétele. Magyar gyomkutatás és<br />

technológia. III (2) 21-29. p.<br />

LEHOCZKY É. és BOROSNÉ N. A. (2002): Echinochloa crus-galli (L.) P. B. és a kukorica<br />

korai kompetíciójának hatása I. Magyar gyomkutatás és technológia. III (1) 13-21.<br />

p.<br />

LEHOCZKY É. és REISINGER P. (2002): Precíziós eljárások alkalmazása kompetíciós<br />

vizsgálatoknál. Magyar gyomkutatás és technológia. III (2) 49-58. p.<br />

MARSHALL, E. J. P. (2002): The role of weeds in supporting biological diversity within<br />

crop fields. 12 th EWRS Symposium, Arnhem-Wageningen. 50-51 p.<br />

MCDONALD, G. K. (2003): Competitiveness against grass weeds in field pea genotypes.<br />

Weed res. 43 (1) 48-58. p.<br />

MUNIER-JOLAIN, N. M, CHAUVEL, B., GASQUEZ, J. (2002): Long-term modelling of<br />

weed control strategies: analysis of threshold-based options for weed species with<br />

contrasted competitive abilities. Weed Res. 42 (2) 107-122. p.<br />

NAGY I. és LEHOCZKY É. (2002): Herbicid választék Magyarországon napjainkban.<br />

Magyar gyomkutatás és technológia. III (2) 59-69. p.<br />

NAGY J. (1995): A talajmővelés, a mőtrágyázás, a növényszám és az öntözés hatásának<br />

értékelése a kukorica (Zea mays L.) termésére. Növénytermelés 44 (3) 251-260. p.<br />

NAGY J. (1996a): A talajmővelés és a mőtrágyázás hatásának értékelése a kukorica (Zea<br />

mays L.) termésére. Agrokémia és talajtan 45 (1-2). 113-124. p.<br />

NAGY J. (1996b): A mőtrágyázás és a talajmővelés kölcsönhatása a kukoricatermesztésben.<br />

Növénytermelés 45 (3). 297-305. p.<br />

NAGY Z. (2003): A gyengébb talajok alternatív növénye, a mustár. Gyakorlati Agrofórum<br />

14 (3) 60-62. p.<br />

NAKOVA, R. (2003): Study of the competition between wheat and Papaver rhoeas. EWRS<br />

7 th Mediterranean Symposium, Adana, 2003. máj. 6-9. 125-126. p.<br />

NÉMETH I., NAGY B., DORNER Z. (2003): A zöldtrágyanövények hatása a<br />

gyomosodásra. Növénytermelés 52 (5). 469-596. p.<br />

NÉMETH I., SÁRFALVI B. (1998): Gyomfelvételezési módszerek értékelése összehasonlító<br />

vizsgálatok alapján. Növényvédelem, 34 (1) 15-22. p.<br />

PEPÓ P. (2001): A kutatás és innováció szerepe a növénytermesztés fejlesztésében In.<br />

JÁVOR A. és SZEMÁN L. (Szerk.): Innováció, a tudomány és a gyakorlat egysége<br />

az ezredforduló agráriumában. Gödöllı, 2001. máj. 17-18. 19-23. p.<br />

PEPÓ P. (2002): Az ıszi búza fajtaspecifikus tápanyagellátása csernozjom talajon. In PEPÓ<br />

P. és JOLÁNKAI M. (Szerk.): II. Növénytermesztési Tud. Nap. „Integrációs<br />

feladatok a hazai növénytermesztésben” Budapest: MTA, 105-110. . p.<br />

26


Soil management and tillage possibilities in weed control<br />

PETRÓCZI I. (1999): Az Integrált Növényvédıszer-használati Rendszer (IPM) fıbb<br />

jellemzıi és követelményrendszere In: KOVÁCS J. (Szerk.): A növényvédelem<br />

integrált környezetbarát fejlesztésének lehetıségei. Budapest: MTA, 190. p.<br />

PERCZE A. (2002): A mővelési rendszerek hatása a talajállapotra és a gyomosodásra ıszi<br />

búzában. Doktori (Ph.D.) értekezés, Gödöllı, 116. p.<br />

PRISZTER SZ. (1998): Növényneveink. Budapest: Mezıgazda Kiadó, 547. p.<br />

PUMMER L., LADÁNYI E., HOLLÓ S. (1997): A vetésforgó hatása a kukorica termésére<br />

különbözı tápanyagszinteken szántóföldi tartamkísérletben. Növénytermelés 46 (6)<br />

593-602. p.<br />

RÁTÜSSZENTENEK, HVG, 2003. jún. 21.<br />

REISINGER P. et al. (2003): Gyomnövényzet vizsgálatok hántott és hántatlan tarlón.<br />

Magyar gyomkutatás és technológia 4 (1) 31-44. p.<br />

RUZSÁNYI L. (1996): Az aszály hatása és enyhítésének lehetıségei a növénytermesztésben.<br />

5-66. p. In. CSELİTEI L. és HARNOS ZS. (Szerk.): Éghajlat, idıjárás, aszály.<br />

Budapest: Akaprint, .p.<br />

RUZSÁNYI L. (2000): Hidrometeorológiai szélsıségek növénytermesztési értékelése. 145-<br />

159. p. In. NAGY J. és PEPÓ P.(Szerk.): Talaj, növény és környezet kölcsönhatásai.<br />

DE ATC, Debrecen.<br />

RUZSÁNYI L. (2001): A növénytermesztés és a talaj vízháztartásának kapcsolatai. In.<br />

JÁVOR A. és SZEMÁN L. (Szerk.): Innováció, a tudomány és a gyakorlat egysége<br />

az ezredforduló agráriumában. 40-44. p.<br />

SÁRVÁRI M. (1995): A kukoricahibridek termıképessége és trágyareakciója réti talajon.<br />

Növénytermelés 44 (2) 179-190. p.<br />

SCHMIDT R., SZAKÁL P., KEREKES G., BENE L. (1998): Soil compaction studies in<br />

tramline sugarbeet cultivation. Proc. Conference on Soil condition and crop<br />

production. Gödöllı 108-111. p.<br />

SHRESTHA, A., KNEZEVIC, S. Z., ROY, R. C., BALL-COELHO, B. R., SWANTON, C.<br />

J. (2002): Effect of tillage, cover crop and crop rotation on the composition of weed<br />

flora in a sandy soil. Weed Research 42. 76-78. p.<br />

SIPOS G. (1972): Földmőveléstan. 5. átdolg. kiadás, Budapest: Mezıgazdasági Kiadó 329.<br />

p.<br />

SIPOS S. (1978): Földmővelési rendszerek 304-317. p. In: LİRINCZ J. (Szerk.)<br />

Földmőveléstan. Budapest: Mezıgazdasági Kiadó, 330 p.<br />

SIPOS S. és HEGEDŐS I. (1982): A talajfizikai állapot, valamint a mőtrágyázás és a<br />

kukorica termése közötti összefüggések. Agrokémia és talajtan 31 (3-4) 339-356. p.<br />

SOLYMOSI P. (1999): Gyomszabályozás növényekbıl származó természetes vegyületekkel.<br />

57-78. p. In. KOVÁCS J. (Szerk.): A növényvédelem integrált környezetbarát<br />

fejlesztésének lehetıségei. Budapest: MTA Agrártudományok Osztálya, 190 p.<br />

STREIT, B., RIEGER, S. B., STAMP, P., RICHTER, W. (2003): Weed populations in<br />

winter wheat as affected by crop sequence, intensity of tillage and time of herbicide<br />

application in a cool and humid climate. Weed res. 43 (1) 20-32. p.<br />

SURÁNYI J. (1952): A szántóföldi kettıstermesztés módszerei és növényei. Budapest:<br />

Mezıgazdasági Kiadó, 110. p.<br />

SZALAI T. (1996): Növénytermesztési rendszerek. p. 265-298 In: BIRKÁS M. (Szerk.):<br />

Földmővelés és földhasználat. Egyetemi jegyzet, Gödöllı 314. p.<br />

SZALAI T. (1999): A talajmővelési és növénytermesztési rendszerek néhány agronómiai<br />

összefüggése fenntartható földhasználat kialakításához. Doktori (Ph.D) értekezés,<br />

Gödöllı, 121. p<br />

SZENTEY L. (2000a): Az országos gyommentesítési program végrehajtásáról.<br />

Növényvédelem, 36 (6) 333-335. p.<br />

27


Aniko Farkas<br />

SZENTEY L. (2000b): İszi búza. In: HUNYADI K., BÉRES I., KAZINCZI G. (Szerk.):<br />

Gyomnövények, gyomirtás, gyombiológia, Budapest: Mezıgazda Kiadó, p. 477.<br />

SZENTPÉTERY ZS., JOLÁNKAI M., HEGEDŐS Z., KASSAI K. (2002). Agrotechnikai<br />

tényezık hatása a búza termésére és beltartalmára. In BENKÓNÉ PONGÓ D és<br />

TÓTH L. (Szerk.): MTA-AMB 26. Kutatási és Fejlesztési Tanácskozás, Gödöllı 31.<br />

p.<br />

SZİKE L. (2001): A melegigényes gyomfajok gyors terjedése és a klímaváltozás<br />

összefüggései. Növényvédelem, 37 (1) 10-12. p.<br />

STEFANOVITS P. (1992): Talajtan, 3. kiadás, Budapest: Mezıgazda Kiadó, p.379.<br />

SVÁB J. (1981): Biometriai módszerek a kutatásban. 3. átdolg. és bıv. kiadás. Budapest:<br />

Mezıgazdasági Kiadó, 521. p.<br />

TÓTH Á. és SPILÁK K. (1998): A IV. Országos Gyomfelvételezés tapasztalatai.<br />

Növényvédelmi Fórum, Keszthely, 49 p.<br />

TUESCA, D., PURICELLI, E., PAPA, J. C. (2001): A long-term study of weed flora shifts<br />

in different tillage systems. Weed res. 41 (4) 369-382 p.<br />

TURSUN, N. és ÖZER, Z. (2003): Research on yield lost of corn, french bean and mixed<br />

french bean – corn caused by weeds. EWRS 7 th Mediterranean Symposium, Adana,<br />

2003. máj. 6-9. 123-124. p.<br />

UJVÁROSI M. (1954): A szántóföldi asszociációk új értelmezése. Botanikai közlemények 44<br />

(3-4) p.183-192<br />

UJVÁROSI M. (1957): Gyomnövények, gyomirtás. Budapest: Mezıgazdasági Kiadó, 789.<br />

p.<br />

UJVÁROSI M. és VINCZE M. (1979): Gyomnövények, gyomirtás. Gödöllı: Egy. jegyzet.<br />

162. p.<br />

VAN DELDEN, A., LOTZ, L. A. P., BASTIAANS, L., FRANKE, A. C., SMID, H. G., R.<br />

M. W. GROENEVELD, KROPFF, M. J. (2002): The influence of nitrogen supply<br />

on the ability of wheat and poteto to supress Stellaria media growth and<br />

reproduction. Weed res. 42 (6) 429-445. p.<br />

VARGA P. (2002): Herbicid- és tápanyagstressz hatása a gyomnövények és a kukorica<br />

produktivitására. Doktori (PhD) értekezés, Keszthely 228 p.<br />

VÁRALLYAY GY. (1997): A mezıgazdaság szerkezetátalakításának veszélyei talaj- és<br />

vízkészleteinkre In: A talajok állapota és a növénytermesztés esélyei. Workshop,<br />

GATE,. aug. 29. 4-5. p.<br />

VÁRALLYAY GY. (1999): A talajfizika gyakorlati alkalmazásai a fenntartható<br />

talajhasználatban. Gyakorlati Agrofórum 9 (7) 4-7. p.<br />

VÁRALLYAY GY. (2003): Növényi tápanyagellátás és a talaj vízgazdálkodása. In:<br />

CSORBA ZS., JOLÁNKAI P., SZÖLLİSI G. (Szerk.): III. Növénytermesztési Tud.<br />

Nap. „A szántóföldi növények tápanyagellátása.” Budapest: Akaprint Kiadó. 7-15.<br />

p.<br />

VINCZE M. (1984): Szántóföldi gyomnövények. Gödöllı: Tankönyvkieg. jegyzet, 115 p.<br />

VIRÁG Á. (1973): Vegyszeres gyomirtás p. 181-184. In: UJVÁROSI M. (1973): Gyomirtás.<br />

Budapest: Mezıgazdasági Kiadó. 288 p.<br />

YAACOBY, T. és RUBIN, B (2002): New invasive weeds in Israel. 12 th EWRS Symposium,<br />

Wageningen-Arnhem 342-343. p.<br />

FARKASNÉ SZERLETICS A. 2002. Mővelési rendszerek értékelése gyomszabályozási<br />

szempontból gödöllıi barna erdıtalajon. Növénytermelés, 51. 5. 513-528.<br />

FARKASNÉ SZERLETICS A. 2003. A parlagfő (Ambrosia artemisiifolia L.) jelenléte és<br />

borítási százalékának változása különbözı mővelési eljárások hatására.<br />

Növényvédelem, 39. 7. 303-311.<br />

FARKAS, ANIKÓ, 2002. Effect of cultivation systems on the weed cover percentage on<br />

gödöllı brown forest soils. Herbologia, Sarajevo, 3. 1. 25-40.<br />

28


Soil management and tillage possibilities in weed control<br />

FARKAS, ANIKÓ: 2002. Presence and cover percentage of Ambrosia artemisiifolia L. as<br />

the result of different cultivation methods. Herbologia, Sarajevo, 3. 1. 13-24.<br />

FARKAS, A. 2000. Technische Möglichkeiten zur Unkrautbekämpfung im integrierten<br />

Pflanzenschutz.. 52. Deutsche Pflanzenschutztagung, Freising-Weihenstephan,<br />

2000. okt. 9-12. Abstracts, p. 486.<br />

FARKAS, A., PERCZE, A., GYURICZA, C., 2002. Effect of different Soil Tillage practices<br />

on the Weed Flora on sandy loam Soil (Chromic Luvisol). 12 th EWRS Symposium.<br />

Arnhem, 2002. jún. 24-27. Proceedings, pp. 28-29.<br />

FARKAS, A. 2002. Verwendung von Bodenbearbeitungsverfahren gegen Verunkrautung im<br />

Mais. 53. Deutsche Pflanzenschutztagung, Bonn, 2002. szept. 16-19. Abstracts, pp.<br />

482-483.<br />

FARKAS, A., FEJİS, Z. D. 2003. Effect of different soil tillage and fertilisation levels on<br />

soil cover of Ambrosia artemisiifolia. 7 th EWRS Mediterranean Symposium Adana.<br />

2003. máj. 6-9. Proceedings, pp. 21-22.<br />

FARKAS, A. 2003. Effect of different soil tillage and fertilization level on soil cover of<br />

Ambrosia artemisiifolia L. Proceedings of the 2 nd Weed Conference in Sarajevo.<br />

2003. jún. 6-7. in Herbologia 4. 1. pp. 85-89.<br />

FARKAS, A. 2003. Effect of different soil tillage and fertilization level on weed cover in<br />

maize. Proceedings of the 2 nd Weed Conference in Sarajevo. 2003. jún. 6-7. in<br />

Herbologia 4. 1. pp. 157-162.<br />

FARKAS, A. 2004. Die Wichtigkeit der Nährstoffversorgung gegen des Unkrautes Ambrosia<br />

artemisiifolia (L.) 22. Deutsche Arbeitsbesprechung über Fragen der<br />

Unkrautbiologie und –bekämpfung Stuttgart-Hohenheim, Euroforum 2. - 4. März,<br />

in Zeitschrift für Pflanzenkrankheiten und Pflanzenschutz, Sonderheft XIX. p. 279-<br />

284.<br />

DORNER, Z., NÉMETH, I., FARKAS, A. The effect of extensive farming on the weed<br />

composition of cereals between 2000 and 2003 in Hungary. 22. Deutsche<br />

Arbeitsbesprechung über Fragen der Unkrautbiologie und –bekämpfung Stuttgart-<br />

Hohenheim, Euroforum 2. - 4. März, in Zeitschrift für Pflanzenkrankheiten und<br />

Pflanzenschutz, Sonderheft XIX. p. 113-117.<br />

FARKAS, A. 2004. The effect of cultivation methods on soil compaction and weediness in<br />

oil seed rape grown as a cover crop in Gödöllı. VII. Kongress on weeds. Serbia,<br />

Palic, 7-11. Jun. 2004. Acta herbologica, Beograd, 13. 2. p. 379-384.<br />

FARKAS I-NÉ, VINCZE M., KASSAI M. K. 2000. Mővelés, talajállapot és gyomosodás<br />

összefüggései (Effect of Soil Tillage on Soil Condition and Weed Infestation).<br />

MTA AMB. 24. Kut. és Fejl. Tanácskozás, Gödöllı, jan.18-19. Kiadvány (szerk.<br />

Tóth L., Benkóné Pongó D.), 2. köt. pp. 15-19.<br />

FARKAS I-NÉ, VINCZE M., PERCZE A., KASSAI M. K. 2001. A gyomszabályozás<br />

agrotech-nikai lehetıségeinek vizsgálata gödöllıi termıhelyen (Examination of<br />

Agricultural Possibilities of Weed Control at Gödöllı Site), MTA AMB. 25. Kut. és<br />

Fejl. Tanácskozás, Gödöllı, jan.23-24. Kiadvány (szerk. Tóth L., Benkóné Pongó<br />

D.), 2. köt. pp. 106-110.<br />

FARKAS A., PERCZE A., VINCZE M. 2001. A SEGAL 65WG hatása különbözı<br />

talajhasználati rendszerekben. 47. Növényvédelmi Tudományos Napok, Budapest,<br />

MTA, 2001. febr. 27-28., Kiadvány (szerk. Kuroli G., Balázs K., Szemessy Á.) p.<br />

125.<br />

FARKAS A. 2001. Környezetbarát növényvédelemért-agrotechnikai gyomirtással. VII.<br />

Ifjúsági Tudományos Fórum. Keszthely, 2001. márc. 29. CD-kiadvány<br />

FARKASNÉ SZERLETICS A., Ujj A. 2001. Talajhasználati tartamkísérletek a környezetgazdálkodás<br />

szolgálatában. MTA MTB II. Növénytermesztési Tudományos Nap<br />

29


Aniko Farkas<br />

„Integrációs feladatok a hazai növénytermesztésben” Proceedings (szerk. Pepó P.,<br />

Jolánkai M.), pp. 180-184.<br />

FARKASNÉ SZ. A., GYURICZA Cs. 2001. Különbözı mővelési eljárások<br />

gyomviszonyokra gyakorolt hatásának összehasonlító értékelése gödöllıi<br />

termıhelyen. XLIII. Georgikon Napok, Keszthely, 2001. szept.20-21. Kiadvány 2.<br />

kötet, pp. 845-849.<br />

FARKASNÉ SZERLETICS A. 2002. A parlagfő (Ambrosia elatior) jelenléte és borítási %-<br />

ának változása különbözı mővelési eljárások hatására. 48. Növényvédelmi<br />

Tudományos Napok, Budapest, MTA, 2002. márc. 6-7., Kiadvány (szerk. Kuroli<br />

G., Balázs K., Szemessy Á.) p. 110.<br />

FARKASNÉ SZERLETICS A. 2002. Gyomborítás változása különbözı talajmővelési<br />

eljárások hatására gödöllıi termıhelyen (Effect of Different Soil Tillage on the<br />

Weed Flora on Sandy Loam Soil). Innováció, a tudomány és a gyakorlat egysége az<br />

ezredforduló agráriumában. SZIE–DE ATC, Debrecen, 2002. április 11-12.<br />

Kiadvány „Növénytermesztés” (szerk. Jávor A., Sárvári M.), pp.312-317.<br />

FARKASNÉ SZERLETICS A., DORNERNÉ FEJİS Z., NÉMETH I. 2002. A parlagfő<br />

borításának alakulása eltérı tápanyagmennyiségek hatására gödöllıi termıhelyen.<br />

HWRS Konferencia, 2002. nov. 14.<br />

DORNERNÉ FEJİS Z., BLASKÓ D., FARKASNÉ SZERLETICS A., NÉMETH I. 2002.<br />

Vetésforgókísérlet biotermesztésben. HWRS Konferencia 2002. nov. 14.<br />

FARKASNÉ SZERLETICS A. 2003. A parlagfő (Ambrosia artemisiifolia L.) borításának<br />

változása eltérı tápanyagmennyiségek és talajmővelés hatására. XIII. Keszthelyi<br />

Növényvédelmi Fórum, 2003. jan. 29-31. pp. 10-13.<br />

FARKASNÉ SZERLETICS A. 2003. A tápanyagellátás jelentısége a gyomszabályozásban<br />

(Importance of nutrition in the weed management). III. Növénytermesztési<br />

Tudományos Nap, Gödöllı, 2003. máj. 15. Kiadvány (szerk. Csorba Zs., Jolánkai<br />

P., Szöllısi G.), pp. 49-53.<br />

FARKASNÉ SZERLETICS A. 2004. Mővelés okozta talajállapotváltozás és gyomosodás<br />

összefüggése száraz évjáratban, ıszi búzában. MTA-AMB Kutatási és Fejlesztési<br />

tanácskozás, Gödöllı, 2004. jan. 20-21. 3. kötet, p. 82-86.<br />

FARKASNÉ SZERLETICS A. 2004. Mővelés, talajállapot és gyomosodás összefüggései<br />

olajretek köztes növényben, gödöllıi termıhelyen. Keszthelyi Növényvédelmi<br />

Fórum, 2004. jan. 28-30. p. 9-11.<br />

FARKASNÉ SZERLETICS A. 2004. Kettıs termesztés, köztes termesztés, köztes növény<br />

elnevezések és használatuk. XLVI. Georgikon Napok, Keszthely, 2004. szept. 16-<br />

17. CD kiadvány, ISBN 963 9096 962<br />

FARKASNÉ SZERLETICS A. 2004. Olajretek alkalmazása védınövényként gödöllıi barna<br />

erdıtalajon. XLVI. Georgikon Napok, Keszthely, 2004. szept. 16-17. CD kiadvány,<br />

ISBN 963 9096 962<br />

30


Herbologia Vol. 7, No. 1, 2006.<br />

RESISTANCE STUDY OF AMARANTHUS RETROFLEXUS L. SPECIES<br />

POPULATION TO THE HERBICIDE IMAZETHAPYR<br />

Branko Konstantinović, Maja Meseldžija, Dragana Šunjka<br />

Faculty of Agriculture, Trg Dositeja Obradovica 8, Novi Sad, Serbia and Montenegro<br />

E-mail: maja@polj.ns.ac.yu<br />

Abstract<br />

Use of herbicides and lack of alternative methods of weed control in<br />

conditions of intensive agricultural production have created convenient<br />

environment for herbicide resistant weed species development. Permanent<br />

use of herbicides belonging to the group ALS inhibitors, especially<br />

imidazolinones and sulfonylureas, led to resistance occurrence of weed<br />

species to this herbicide group.<br />

During two years (2005-2006) resistance of weed species Amaranthus<br />

retroflexus L. to ALS inhibitors was studied. From different localities in<br />

Vojvodina, i.e. Krivaja, Kikinda and Becej, with a long history of ALS<br />

inhibitors use in weed control, seed of plants for which there exist possibility<br />

of resistance occurrence to herbicide imazethapyr was collected. Studies were<br />

performed by two methodic procedures, by Petri dish assays (Clay and<br />

Underwood, 1990) and by whole plant studies (Moss, 1995). In the trials, as a<br />

susceptible standard herbicide free population of Amaranthus retroflexus L.<br />

from ruderal sites that was used. Results of the assay are given in the reaction<br />

curve, and resistance index is determined in regard to the susceptible referent<br />

population.<br />

By comparative analysis, resistance occurrence of weed biotype<br />

Amaranthus retroflexus L. from the locality Krivaja has been established,<br />

whereas populations from localities Becej and Kikinda remained susceptible<br />

to the mode of action of the studied herbicide belonging to the ALS inhibitor<br />

group.<br />

Key words: Amaranthus retroflexus L., ALS inhibitors, imazethapyr, herbicide resistance.<br />

Introduction<br />

Nowadays herbicides belonging to the group of ALS inhibitors such<br />

as sulfonylureas, imidazolinones, triazolopyirimidines, pyrimidinylthiobenzoates<br />

and sulfonylamino-carbonil-triazoles represent extremely<br />

significant herbicides for weed control in many crops (Wagner et al., 2002).<br />

Their permanent use during three years period, or longer, resulted in<br />

evolution of ALS resistant biotypes in the world (Thill et al., 1991; Powles


B. Konstantinović et al.<br />

and Shaner, 2001; Rashid et al., 2003; Rubin et al., 2004) and in Vojvodina<br />

(Konstantinovic et al., 2003a; 2003b; 2003c).<br />

Primary action target site of these herbicides is the enzyme of<br />

acetolactate synthase that participate in biosynthesis of amino acids<br />

isoleucine, leucine and valine (Ray, 1984; Gerwick et al., 1990; Takahashi et<br />

al., 1991, Babczinski, 2002), and resistance develops as the consequence of<br />

the single mutation points that ALS structure makes less susceptible to<br />

herbicides.<br />

After five years of permanent use of the ALS inhibiting herbicides in<br />

monoculture of Triticum aestivum L. crop (Mallory-Smith et al., 1990) in<br />

1987 the first weed resistant to ALS inhibitors, Lactuca serriola (Smith and<br />

Cairns, 2001) was determined. Stellaria media (L.) Vill (Kudsk et al., 1995)<br />

was the first weed species resistant to Sulfonylureas found in Europe. Until<br />

now, resistance has been described for 93 weed biotypes from all over the<br />

world, and this number is constantly increasing (Heap, 2006). In regard to the<br />

other herbicide groups, weed biotypes resistant to herbicide that inhibit ALS<br />

enzyme are the most numerous.<br />

In the paper the occurrence of herbicide resistance development of the<br />

weed species Amaranthus retroflexus L. from various localities in Vojvodina<br />

to herbicide imazethapyr from the chemical family Imidazolinones. On the<br />

studied localities Imidazolinones had long history of use in weed control.<br />

Material and methods<br />

During 2005 and 2006 study of weed species Amaranthus retroflexus<br />

L. resistance to imazethapyr was performed according to the methods applied<br />

in laboratory conditions and climatic chamber (Clay and Underwood, 1990;<br />

Moss, 1995). In bioassays a range of imazethapyr rates were used, e.g. 0 ,<br />

0.04 , 0.08 , 0.10 , 0.15 , 0.20 and 0.40 kg a.i./l. During assays seedling<br />

epicotyls and hypocotyls length, stem height, foliage fresh weights and seed<br />

germination and shooting were measured. Statistical data processing was<br />

performed by variance analysis (ANOVA), and significant difference was<br />

evaluated by t-test (Hadzivukovic, 1991). Resistance can be determined only<br />

if there are statistically significant differences between the studied population<br />

and the susceptible standard and if resistant population can not be controlled<br />

by herbicide rate that is efficient in control of the susceptible one (Beckie et<br />

all., 2000). Results of the measured parameters – epicotyls and hypocotyls<br />

length, stem height and germination percentage and shooting are presented as<br />

a quantity-response curve, while values of the foliage fresh weight and<br />

resistance indices were given in a table.<br />

32


Resistance study of A. retroflexus L. species population to the herbicide imazethapyr<br />

Results<br />

Statistical analysis results of epicotyls and hypocotyls shoots length,<br />

stem height and foliage fresh weight (t test) are given in Table 1. Analysis of<br />

imazethapyr effect to the measured biological parameters of the species<br />

Amaranthus retroflexus L. showed that there are statistically significant<br />

differences (p< 0.05) between values of the measured parameters of<br />

populations from locality Krivaja and population used as a susceptible<br />

standard (S).<br />

Tab. 1. Significant difference between Amaranthus retroflexus L. populations<br />

from the studied localities and susceptible standard treated by imazethapyr<br />

Localities<br />

Parameters<br />

Hypocotyl<br />

length<br />

Stem<br />

height<br />

Foliage<br />

fresh<br />

weight<br />

Epicotyl<br />

length<br />

Krivaja - S * * * *<br />

Kikinda - S N.Z. N.Z. N.Z. N.Z.<br />

Bečej - S N.Z. N.Z. N.Z. N.Z.<br />

p


B. Konstantinović et al.<br />

Amaranthus retroflexus L. seed germination<br />

Effect of the herbicide imazethapyr to seed germination of weed<br />

species Amaranthus retroflexus L. is given in the Graph 1. The highest<br />

percentage of germinated seeds was established for population from locality<br />

Krivaja. Population of the susceptible standard and localities Kikinda and<br />

Becej had very low germination capability with all applied herbicide rates.<br />

duzina epikotila (mm)<br />

epycotyl length (mm)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0<br />

0.04<br />

0.08<br />

0.1<br />

0.15<br />

0.2<br />

0.4<br />

kg a.m. imazetapir/l<br />

kg. a.i. imazethapyr/l<br />

Krivaja<br />

Kikinda<br />

Bečej<br />

Ruderalno<br />

staniste/ruderal<br />

site<br />

Fig. 2. Effect of the herbicide imazethapyr to Amaranthus retroflexus<br />

L. epycotyl seedlings length<br />

34


Resistance study of A. retroflexus L. species population to the herbicide imazethapyr<br />

duzina hipokotila (mm)<br />

hypocotyl length (mm)<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 0.04 0.08 0.1 0.15 0.2 0.4<br />

kg a.m. imazetapir/l<br />

kg a.i. imazethapyr/l<br />

Krivaja<br />

Kikinda<br />

Bečej<br />

Ruderalno<br />

staniste/ruderal site<br />

Fig. 3. Effect of the herbicide imazethapyr to Amaranthus retroflexus<br />

L. hypocotyl seedlings length<br />

By statistical analysis (Table 1) of the values for epicotyls and<br />

hypocotyls length, significant difference was determined for populations<br />

from locality Krivaja in regard to the other studied localities and population<br />

of the susceptible standard. Populations values from localities Kikinda and<br />

Becej did not show significant differences from those obtained for the<br />

susceptible standard. Values for epicotyls and hypocotyls length are<br />

presented in Graphs 2 and 3.<br />

procenat nicanja<br />

% of shooting<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 0.04 0.08 0.1 0.15 0.2 0.4<br />

kg a.m. imazetapir/l<br />

kg a.i. imazethapyr/l<br />

Krivaja<br />

Kikinda<br />

Bečej<br />

Ruderalno<br />

staniste/ruderal site<br />

Fig. 4. Effect of the herbicide imazethapyr to shooting of<br />

Amaranthus retroflexus L. plants<br />

35


B. Konstantinović et al.<br />

Effect of the herbicide imazethapyr to shooting of plants is given in<br />

Graph 1. Percentage of emerged plants from the site Krivaja was the highest.<br />

visina stabla (mm)<br />

stem height (mm)<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 0.04 0.08 0.1 0.15 0.2 0.4<br />

kg a.m. imazetapir/l<br />

kg a.i. imazethapyr/l<br />

Krivaja<br />

Kikinda<br />

Bečej<br />

Ruderalno<br />

staniste/ruderal site<br />

Fig. 5. Effect of the herbicide imazethapyr to the height Amaranthus<br />

retroflexus L. stems.<br />

Values of the studied parameters of Amaranthus retroflexus L.<br />

biotype plants from locality Krivaja showed statistical significant difference<br />

(p< 0.05) in regard to the susceptible standard and other studied localities.<br />

Between biotypes from localities Kikinda and Becej there were no<br />

statistically significant differences, nor there were differences in regard to the<br />

susceptible standard (Graph 4).<br />

Resistance level based upon foliage fresh weight was determined<br />

according to the scale by Moss et al (1999) that imply several possible<br />

resistance levels of the studied population (Table 2).<br />

Tab. 2. Resistance level of the studied populations of Amaranthus<br />

retroflexus L. determined according to the scale by Moss, based upon<br />

the fresh weight.<br />

Locality Resistance level<br />

Krivaja 4*<br />

Kikinda 1*<br />

Bečej<br />

S<br />

S - osetljivost na primenjeni herbicid/susceptibility to the applied herbicide<br />

1* – rana indikacija rezistentnosti, mogućnost da je došlo do redukcije delovanja<br />

herbicida/early indication of the resistance, suspected herbicide reduction<br />

36


Resistance study of A. retroflexus L. species population to the herbicide imazethapyr<br />

2*/3* - potvrñena rezistentnost, mogućnost da je došlo i do redukcije delovanja<br />

herbicida/confirmed resistance, suspected herbicide reduction<br />

5*/4* - potvrñena rezistentnost, mala verovatnoća lošeg delovanja herbicidacomfirmed<br />

resistance/reduced likelihood of low herbicide action.<br />

Tab. 3. Resistance level of the studied populations of Amaranthus<br />

retroflexus L. given in resistance index<br />

Resistance index<br />

Locality epicotyl hypocotyl stem height foliage fresh weight<br />

Krivaja 1.8 1.61 1.0 2.27<br />

Kikinda 1.0 0.8 0.9 0.53<br />

Bečej 1.0 1.0 1.0 0.99<br />

The highest values of resistance index of the measured parameters<br />

were determined for population from the locality Krivaja (1.0 – 2.27). The<br />

values of resistance index for population of Amaranthus retroflexus L. from<br />

localities Kikinda and Becej were significantly lower, i.e. less than 1.<br />

Discussion<br />

Determined significant differences in germination (Fig. 1) and values<br />

for epicotyls and hypocotyls length (Fig. 2 and 3) between biotype from<br />

locality Krivaja and susceptible standard, as well as other studied localities<br />

implied high resistance of this biotype to the herbicide imazethapyr. It was<br />

also determined that decay of seedlings from localities Kikinda and Becej<br />

occurred at rates above 0.08 kg a.i. imazethapyr/l, while seedlings from<br />

localities Krivaja at rate of 0.40 kg a.i. imazethapyr/l remained well<br />

developed. Seedlings from ruderal site, used as the susceptible standard<br />

completely decayed at a rate of 0.04 kg. a.i. imazethapyr/l.<br />

In regard to susceptible standard and localities Kikinda and Becej,<br />

increased rates of the applied herbicide caused the lowest reduction in<br />

epicotyls and hypocotyls length of weed species Amaranthus retroflexus L.<br />

biotype from locality Krivaja.<br />

Imazethapyr effect to weed species Amaranthus retroflexus L.<br />

biotypes evinced in differences in plant emergence and stem height (Fig. 4<br />

and 5). In all applied imazethapyr rates, the highest percentage of emerged<br />

plants (100%), as well as the highest values for the stem height was<br />

determined for biotype from locality Krivaja. Decay of plants from this<br />

locality did not occur even with applied rate of 0.40 kg a.i. imazethapyr/l.<br />

Based upon measured biological parameters, samples of the susceptible<br />

standard, as well as plants from localities Kikinda and Becej, remained<br />

37


B. Konstantinović et al.<br />

susceptible to herbicide imazethapyr action, for decay occurred after<br />

application of the second rate of 0.08 kg a.i. imazethapyr/l.<br />

Based upon resistance level obtained by foliage fresh weight<br />

measurement (Table 2), resistance was confirmed for populations from<br />

locality Krivaja (4*), with a slight probability that herbicide showed low<br />

efficiency. For biotype from locality Kikinda an early indication of resistance<br />

was determined, but with a possibility of herbicide efficiency reduction.<br />

Population from locality Becej remained susceptible to the applied herbicide.<br />

Resistance index of all measured parameters (Table 3) suggests that<br />

biotype from the locality Krivaja acquired the highest resistance to the<br />

applied rates of the herbicide imazethapir. Populations from locality Becej<br />

(IR = 0.99 - 1) and Kikinda (IR = 0.53 – 1) showed high susceptibility to<br />

herbicide imazethapyr.<br />

Conclusion<br />

Based upon results of the resistance studies of different populations<br />

of the weed species Amaranthus retroflexus L. to herbicide imazethapyr, it<br />

can be concluded that intensive use of ALS inhibiting herbicides caused<br />

reduced susceptibility of weed species Amaranthus retroflexus L. population<br />

to imazethapyr. The biotype from the site Krivaja showed the highest<br />

resistance to the applied rates of herbicide imazethapyr. Biotype from the<br />

locality Becej showed reduced susceptibility to the applied herbicide (RI =<br />

0.99 – 1.0). The highest susceptibility to the applied rates of the herbicide<br />

imazethapyr was found in the biotype from Kikinda (RI = 0.53 – 1.0).<br />

References<br />

BABCZINSKI, P. (2002): Discovery of the lead structure for propoxycarbazone -sodium (BAY<br />

MKH 6561). Pflanzenschutz-Nachrichten-Bayer 55, p. 5-14.<br />

BECKIE, H.J., HEAP, J.M., SMEDA, R.J., HALL, L.M. (2000): Screening for Herbicide<br />

Resistance in Weeds. Weed Technology, 14, p. 428-445.<br />

CLAY, D.V. & UNDERWOOD, C. (1990): The identification of triazine and paraquat resistant<br />

weed biotypes and their response to other herbicides. Importance and perspectives<br />

on herbicide resistant weeds, Luxemburg. p. 47-55.<br />

GERWICK, C.M., SUBRAMANIAN, M.V. AND LONEY-GALLANT, V.I. (1990): Mechanism of<br />

action of the 1,2,4,-triazolo[1,5-]pyrimidines. Pesticide Science, 29, p. 357-364.<br />

HADŽIVUKOVIĆ, S. (1991): Statistički metodi s primenom u poljoprivrednim i biološkim<br />

istraživanjima. Poljoprivredni fakultet, Novi Sad.<br />

HEAP, I. (2006): International survey of herbicide resistant weeds. Online internet,<br />

www.weedscience.com<br />

KONSANTINOVIĆ, B., MESELDŽIJA, M., ŠUNJKA, D., KONSTANTINOVIĆ, BO. (2003a):<br />

Determination of resistant biotypes of Amaranthus retroflexus L. on ALS inhibitors.<br />

3 rd International Plant Protection Symposium at Debrecen University, Debrecen<br />

Hungary, p. 269-276.<br />

38


Resistance study of A. retroflexus L. species population to the herbicide imazethapyr<br />

KONSTANTINOVIĆ, B., MESELDŽIJA, M., POPOVIĆ, S., KONSTANTINOVIĆ, BO. (2003b): Study<br />

of resistance to ALS inhibitors in the weed species Echinochloa crus-galli L., The<br />

BCPC International Congress, Glasgow, Scotland, UK, Vol. 2, p. 771-775.<br />

KONSTANTINOVIĆ, B., MESELDŽIJA, M., ŠUNJKA, D. (2003c): Ispitivanje rezistentnosti<br />

korovske vrste Echinochloa crus-galli L. na ALS inhibitore, Zbornik referata 2.<br />

savetovanja o korovima, Sarajevo, 6-7 juni 2003.<br />

KUDSK, P., MATHIASSEN, S.K. AND COTTERMAN, J.C. (1995) Sulfonylurea resistance in<br />

Stellaria media (L.) Vill. Weed Research, 35, p. 1924.<br />

MALLORY-SMITH, C.A., THILL, D.C.& DIAL, M.J. (1990): Identification of sulfonilurea<br />

herbicide-resistant prickly lettuce (Lactuca seriola). Weed Technology, 4, p. 163-<br />

168.<br />

MOSS, S.R. (1995): Techiniques for determinig herbicide resistance. Proceedings of the<br />

Brighton Crop Protection Conference-Weeds, p. 547-556.<br />

POWLES, S. B. & SHANER, D.L. EDS. (2001): Herbicide resistance in world grains. CRC<br />

Press, Boca Raton, Florida, USA. p. 308.<br />

RASHID, A., NEWMAN, J.C., O'DONOVAN, J.T., ROBINSON, D., MAURICE, D., POISSON, D. AND<br />

HALL, L.M. (2003): Sulfonylurea herbicide resistance in Sonchus asper biotypes in<br />

Alberta, Canada. Weed Research, Volume 43, p. 214-221.<br />

RAY, T.B.(1984): Site of action of chlorsulfuron. Plant Physiology, 75, p.827-831.<br />

RUBIN, B., TAL, A. AND YASUOR, H. (2004): The significance and impact of herbicide<br />

resistant weeds – a global overveiw. Acta Herbologica, 13, p. 277-288.<br />

SHANER, D.L., ANDERSON, P.C. AND STIDHAM, M.A. (1984): Imidazolinones: potent<br />

inhibitors of acethydroxyacid synthase. Plant Physiology, 76, p. 545-546.<br />

SMIT, J.J. AND CAIRNS, A.L.P. (2001): Resistance of Raphanus raphanistrum in the Republic<br />

of South Africa. Weed Research, 41, p. 41-47.<br />

TAKAHASHI, S., SHIGEMATSU, S. AND MORITA, A. (1991): KIH-2031, a new herbicide for<br />

cotton. Proc. Brighton Crop Prot. Conf., p. 57-62.<br />

THILL, D.C., MALLORY-SMITH, C.A., SAARI, L.L., COTTERMAN, J.C., PRIMIANI, M.M., AND<br />

SALADINI, J.L. (1991) Sulfonylurea herbicide resistant weeds: discovery,<br />

distribution, biology, mechanism and management. In: Herbicide Resistance in<br />

Weeds and Crops (eds JC Caseley, GW Cussans & RK Atkin), 115128.<br />

Butterworth-Heinemann, Oxford.<br />

WAGNER, J., HAAS, H.U. AND HURLE, K. (2002): Identification of ALS inhibitor-resistant<br />

Amaranthus Biotypes using polymerase chain reaction amplification of specific<br />

alleles. Weed Research, 42, p. 208-287.<br />

39


Herbologia Vol. 7, No. 1, 2006.<br />

INFLUENCE OF THE ADJUVANT DESH ON THE EFFICACY AND<br />

SELECTIVITY OF IMAZAMOX 40 a.i.L -1 (PULSAR 40) IN THREE<br />

PERENNIAL LEGUME CROPS<br />

Tsvetanka Dimitrova 1 , Senka Milanova 2<br />

1 Institute of Forage Crops, 5800 Pleven ifc@el-soft.com<br />

2 Institute of Plant Protection, 2230 Kostinbrod protection@infotel.bg<br />

Abstract<br />

During the period 2003-2005 a study was conducted on slightly<br />

leached chernozem with the purpose of studying the influence of the adjuvant<br />

Desh on the efficacy and selectivity of Imazamox 49 a.i.L -1 (Pulsar 40) in<br />

lucerne (Medicago sativa L.), birdsfoot trefoil (Lotus corniculatus L.) and<br />

sainfoin (Onobrychis vicifolia Scop.). It was found that:<br />

The herbicide Imazamox 40 a.i.L -1 (Pulsar 40) at the rate of 20 ml<br />

a.i.ha -1 in combination with the adjuvant Desh – 1000 ml a.i. ha -1 applied in<br />

early growing season of lucerne, birdsfoot trefoil and sainfoin had high<br />

selectivity and herbicidal efficacity reaching 93-97%;<br />

Treatment of swards of perennial legume crops improved their<br />

botanical composition and increased dry biomass productivity 1.4 to 2.8<br />

times.<br />

Key words: Imazamox 40 a.i.L -1 , adjuvant Desh, weeds, productivity, perennial legume<br />

crops.<br />

Introduction<br />

Economic importance of the perennial legume crops including lucerne<br />

(Medicago sativa L.), birdsfoot trefoil (Lotus corniculatus L.) and sainfoin<br />

(Onobrychis vicifolia Scop.) is many-sided. They are valuable for the high<br />

biological value of their forage, as well as in an ecological aspect improving<br />

soil fertility and phytosanitary state.<br />

Weed competition is one of the main factors causing quick thinning<br />

of the swards of these species, reducing and worsening quality of their<br />

production. This circumstance is of decisive importance for weed control<br />

during the period after establishment of the stands of these herbaceous<br />

species.<br />

Although the chemical method of controlling weeds in the swards of the<br />

above-mentioned species is considered efficient by some authors, the studies<br />

are limited in this field (Benkov & Prodanov, 1975; Dimitrova, 1987 and<br />

1995; Lescar Audy, 1971).<br />

The out-of-vegetation treatment of lucerne with Imazethapyr 100


Tsvetanka Dimitrova and Senka Milanova<br />

a.i.L -1 (Speed 10 SL) resulted in 90.8 % biological efficacy, increase of dry<br />

biomass and crude protein yield of 60 % and 2.4 to 2.6 % respectively and<br />

decrease of fibre content of 3.8 to 4.1 % (Dimitrova, 2001). The chemical<br />

weed control proved to be the most efficient in pure lucerne growing for<br />

forage, the net income of its application increasing by 64.8 % (Stoykova &<br />

Dimitrova, 2005).<br />

The long and unilateral use of the same herbicides resulted in<br />

selection of genetically resistant weeds (Nikolova & Konstantinov, 1989;<br />

Beckie et al., 2000; Lee & Owen, 2000). Resistance has been reported for<br />

most herbicide categories and at least for 174 weed species (Heap, 2004).<br />

This circumstance necessitates study of new approaches to using herbicides,<br />

new active substances, optimimization of their doses. Some authors reported<br />

that the adjuvants to the herbicidal solutions led to an increase of their<br />

efficacy and to reduction of the doses (Kudsk & Streibig, 2002; Dogan et al.,<br />

2002).<br />

Imazamox belongs to the imidazolinone class that includes imazapyr,<br />

imazapic, imazethapyr, imazamox and imazametabenz. The herbicides of the<br />

group of imidazolinones and sulphonylureas kill the weeds by inhibiting<br />

acetolactate synthase (ALS). The imidazolinone herbicides possess high<br />

biological efficacy at low application rates and are an attractive alternative<br />

for weed control in lines of spring wheat resistant to the imidazolinone group<br />

(Pozniak et al., 2004), winter wheat (Stougaard et al., 2004). Resistance to<br />

imazamox has also been introduced into cultivated sunflower by traditional<br />

breeding methods (Massinga et al., 2005). Herbicide-resistant rape prevails<br />

on the rape market in Canada with imidazoline-resistant canola (IPI) with<br />

50:50 combination of imazamox and imazethapyr (Karker et al., 2004). Due<br />

to the great biological activity of imazamox it is very important to know the<br />

possibility for dose decrease. The interest for application of reduced rates of<br />

herbicides is in favor of the farmers and environment. The adjuvants can<br />

stimulate the herbicide uptake and provide a possibility for dose reduction<br />

(Mathiassen & Kudsk, 2002; Kieloch & Domaradzki, 2005).<br />

The objective of this study was to investigate the influence of the<br />

adjuvant Desh on the efficacy and selectivity of Imazamox 40 a.i.L -1 (Pulsar<br />

40) in lucerne (Medicago sativa L.), birdsfoot trefoil (Lotus corniculatus L.)<br />

and sainfoin (Onobrychis vicifolia Scop.).<br />

Material and methods<br />

The study was conducted during the period 2003-2005 at the<br />

experimental field of the Institute of Forage Crops in Pleven on slightly<br />

leached chermozem. The variants presented in Table 1 were laid out three<br />

times by years in established stands of lucerne, birdsfoot trefoil and sainfoin.<br />

42


Influence of the adjuvant Desh on the efficacy and selectivity of imazamox 40 a.i. l -1 (Pulsar)<br />

The block method was used with three replications and harvest plot size of 20<br />

m 2 .<br />

Natural background of weed infestation was used. The predominant<br />

weed species were also main weeds in the old swards of the above-mentioned<br />

crops: Capsella bursa pastoris L., Thlaspi arvense L., Stellaria media L.,<br />

Veronica hederifolia L., Anagalis arvensis.<br />

The treatment was made with 400 l ha -1 working solution in early<br />

growing season. The adjuvant Desh was added immediately before the<br />

treatment. The following characteristics were observed: selectivity (after<br />

EWRS scale); degree of weed infestation (by the quantity and quantityweight<br />

method); herbicidal efficacy, %; dry biomass yield mathematically<br />

processed by the method of variance analysis.<br />

During the three study years the results of the observed characteristics<br />

retained their trend among the variants which allowed their presentation on<br />

average for the experimental period.<br />

Results and discussion<br />

The herbicide Imazamox applied alone or with the adjuvant Desh<br />

possesses high selectivity to lucerne, birdsfoot trefoil and sainfoin. It belongs<br />

to the group of wide-spectrum herbicides with phytotoxic action to the annual<br />

mono- and dicotyledonous weeds. Under the trial conditions Capsella bursa<br />

pastoris L. and Thlapsi arvense L. showed high susceptibility; they are also<br />

the main weeds of the old stands of the studied species. Our observations<br />

showed complete killing of the weeds that were at earlier stages of their<br />

development (seedlings and rosette) at the moment of treatment. The weeds<br />

that were at a more advanced stage remained chlorotic and formed no flowerbearing<br />

stems. In view of the circumstance that in these stands a considerable<br />

part of the weeds wintered successfully, the timely treatment at the first<br />

opportunity in spring was a necessary condition of reaching high herbicidal<br />

efficacy.<br />

The herbicidal efficacy (Table 2) with regard to the weed weight in<br />

the standard and Imazethapyr reached 96-98% in different crops. The values<br />

for Imazamox at the rate of 20 ml a.i.ha -1<br />

with the adjuvant Desh at the rate<br />

of 1000 ml ha -1, being within the range of 93-97 %, were the closest to these.<br />

When comparing this efficacy with that for the application of the herbicide<br />

alone it was evident that owing to the adjuvant Desh it was higher by 17 to<br />

19%. The synergistic action of the activating additives was also reported by<br />

other authors who explained it by an increase of the retention and absorption<br />

of the herbicidal solution by the leaves (Borona et al., 2003; Woznica, 2005).<br />

At the lower dose of the adjuvant of 500 ml ha -1<br />

the herbicidal efficacy<br />

43


Tsvetanka Dimitrova and Senka Milanova<br />

increased by 9 to 10% as compared to its application alone. The results<br />

showed a unsatisfactory herbicidal effect (71-79%) when applying the<br />

herbicide alone (V 3 ), as well as when applying it with the adjuvant Desh at<br />

the lower doses (V 6 ).<br />

The removal of the competitive effect of the weeds led in an increase<br />

of the participation of the cultivated components in the swards of the<br />

perennial legume crops and as a result the dry biomass productivity also<br />

increased (Table 3). This increase was 2.4 to 2.8 times in the treated lucerne<br />

stands and 1.1 to 1.4 times in birdsfoot trefoil and sainfoin.<br />

The highest yields of dry biomass, close to those of the weeded check<br />

and the standard, were harvested from the lucerne, birdsfoot trefoil and<br />

sainfoin stands treated with Imazamox + Desh at the rates of 20 + 1000 ml<br />

a.i.ha -1<br />

reaching 4700, 2690 and 6350 kg ha -1 , respectively. The differences<br />

in the absolute values of the dry biomass yields had very good positive<br />

significance.<br />

Tab. 1. Trial variants<br />

Variant* Rate, ml a.i.ha -1<br />

V 1 – Check – zero -<br />

V 2 – Imazethapyr 100 a.i.L -1 (Pulsar 100SL)-standard 40<br />

V 3 – Imazamox 40 a.i.L -1 (Pulsar 40) 20<br />

V 4 – Imazamox 40 a.i.L -1 (Pulsar 40)+Desh (adjuvant) 20+500<br />

V 5 – Imazamox 40 a.i.L -1 (Pulsar 40)+Desh 20+1000<br />

V 6 – Imazamox 40 a.i.L -1 (Pulsar 40)+Desh 16+500<br />

V 7 – Imazamox 40 a.i.L -1 (Pulsar 40)+Desh 16+1000<br />

V 8 – Check – weeded -<br />

*The variants of V 1 to V 8 were laid out in three perennial legume crops: lucerne (Medicago<br />

sativa L.), birdsfoot trefoil (Lotus corniculatus L.) and sainfoin (Onobrychis vicifolia Scop.)<br />

Tab. 2. Efficacy of Imazamox 40 a.i.L -1 (Pulsar 40) in perennial legume crops<br />

Lucerne<br />

Birdsfoot trefoil<br />

Sainfoin<br />

Variant<br />

(Medicago sativa L.) (Lotus corniculatus L.) (Onobrychis vicifolia Scop.)<br />

Weeds/m 2 Weeds/m 2 Weeds/m 2<br />

number weight,g HE*,% number weight,g HE*,% number weight,g HE*,%<br />

V 1 534 655 - 322 679 - 280 590 -<br />

V 2 10 15 98 12 28 96 6 12 98<br />

V 3 143 161 76 97 175 74 73 125 79<br />

V 4 26 94 86 59 115 83 25 67 89<br />

V 5 12 31 95 27 49 93 9 19 97<br />

V 6 144 169 74 103 199 71 81 141 76<br />

V 7 104 101 85 65 129 81 38 78 87<br />

44


Influence of the adjuvant Desh on the efficacy and selectivity of imazamox 40 a.i. l -1 (Pulsar)<br />

HE* - herbicidal efficacy<br />

Tab. 3. Influence of the treatment with Imazamox 40 a.i.L -1 (Pulsar<br />

40) on dry biomass productivity of perennial legume crops<br />

Dry biomass<br />

Variant Lucerne<br />

(Medicago sativa L.)<br />

Birdsfoot trefoil<br />

(Lotus corniculatus L.)<br />

Sainfoin<br />

(Onobrychis vicifolia Scop)<br />

kg/ha -1 %V 1 kg/ha -1 %V 1 kg/ha -1 %V 1<br />

V 1 1680 100 1820 100 4310 100<br />

V 2 4810 286 2770 152 6350 147<br />

V 3 4210 250 2150 118 5010 116<br />

V 4 4450 265 2370 130 5530 128<br />

V 5 4700 280 2690 148 6350 147<br />

V 6 4150 247 2130 117 4920 114<br />

V 7 4330 258 2260 124 5230 121<br />

V 8 4920 293 2860 157 6430 149<br />

GD P 5% 68,7 140,2 135,1<br />

P 1% 95,4 194,7 187,5<br />

P 0,1% 132,7 270,7 260,8<br />

Conclusion<br />

The herbicide Imazamox 40 a.i.ha -1 (Pulsar 40) at the dose of 20 ml<br />

a.i.ha -1 in combination with the adjuvant Desh at 1000 ml a.i.ha<br />

-1 applied in<br />

early growing seasonon of lucerne, birdsfoot trefoil and sainfoin had high<br />

selectivity and herbicidal efficacy reaching 93-97%;<br />

Treatment of swards of perennial legume crops improved their<br />

botanical composition and increased dry biomass productivity 1.4 to 2.8<br />

times.<br />

Reference<br />

BENKOV, B., I.PRODANOV, (1975): In: Chemical weed control, Zemizdat, Sofia.<br />

DIMITROVA, TS., (1987): Influence of weeds and their control on forage yields and sward<br />

botanical composition of pure and mixed birdsfoot trefoil stands, Plant Science,<br />

XXIV, 1, 65-68.<br />

DIMITROVA, TS., (1995): Study of the influence of growing method and weed control on<br />

the degree of weed infestation and seed yield of sainfoin, Contemporary Plant<br />

Protection (Proc. Papers, October 1995), Sofia, 462-466.<br />

DIMITROVA, TS., (2001): Biological study of the herbicide Speed 10SL (Imazethapyr<br />

100g/l) under out-of-vegetation treatment of lucerne (Medicago sativa L.), Proc.<br />

Scientific Works of AU – Plovdiv, vol. XLVI, No.2, 205-208.<br />

NIKOLOVA, V., K.KONSTANTINOV, (1989): Study of compensation changes of weed<br />

associations in vegetable crops and potatoes – a basis for accurate prognosis and<br />

efficient control, Proc. Plant Protection in Aid of Agriculture, Proc. IPP, USWB,<br />

45


Tsvetanka Dimitrova and Senka Milanova<br />

Sofia, 225-245.<br />

STOYKOVA, M., TS.DIMITROVA, (2005): Comparative economic analysis of efficiency<br />

of the plant protection practices used in lucerne forage production, Agricultural<br />

Economics and Management, 50, 4, 54-58.<br />

BECKIE, H. I., HEAP,I..M., SMEAA, R.I., HALL, L. M., (2000): Screening for herbicide<br />

resistance in weeds, Weed Technology, 14 (2) 428-445.<br />

BORONA, VL., V.ZADOROZHNY,T. POSTOLOVSKAY, (2003): The influence of<br />

adjuvant on the efficacy of graminicides in soybeans and nicosulfuron in maize,<br />

Herbologija, vol. 4, № 1, 151-155.<br />

DOGAN, M.N., BOZ,O., ALBAY, F., (2002): Influence of some additives on the efficacy of<br />

nicosulfuron in maiz and fenoxa-prop-P-ethyl in wheat, Proc. 12 th EWRS<br />

Symposium, Wageningen (The Netherlands), 94-95.<br />

HARKER, K., G. CLAYTON, J. O’DONOVAN, R. BLACKSNAW, F. STEVENSON,<br />

(2004). Herbicide timing and rate effects on weed management in three herbicide –<br />

resistant canola systems. Weed Technology, 18, 4, 1006-1012.<br />

HEAP, I., (2004): International Sulvey of herbicide - resistant weed. http: www weed<br />

science. org. Accessed 22 November, 2004.<br />

KIELOCH, R., K. DOMARADZKI (2005): The influence of relative humidity on Anthemis<br />

arvensis and Stellaria media control by tribenuron - methyl used alone and with<br />

adjuvants. 13 th EWRS Symposium, Bari, Italy, CD – ROM.<br />

KUDSK, P., STREIBIG, I.C., \2002|: Herbicides-a double – edged sword Proc. 12 th<br />

EWRS Symposium, Wageningen (The Netherlands), 94-95.<br />

LEE, I.M., OWEN, M.D., (2000): Comparison of acetolactate synthase enzyme inhibition<br />

amond resistant and susceptible Xanthium strumarium biotypes, Weed Science, 48<br />

(3) 286-290.<br />

LESCAR, L., AUDY, (1971): Comptes rendus des journees d’etudes sur les herbicides, t. 4.<br />

COLUMA.<br />

MASSINGA, R., K. AL-KHATIB, P. ST. AMAND, J. MILLER (2005): Relative fitness of<br />

imazamox resistant common sunflower and prairie sunflower. Weed Science, 53, 2,<br />

166-174.<br />

MATHIASSEN, S., PER KUDSK (2002): The influence of adjuvants on the efficacy and<br />

rainfastness of iudosulfuron. 12 th EWRS (European Weed Research Society)<br />

Symposium 2002, Wageningen, 206-207.<br />

POZNIAK, C., F. HOLM, R. HUCL (2004): Field performance of imazamox – resistant<br />

spring wheat. Canadian Journal of Plant Science, 84, 4, 1205-1211.<br />

STOUGAARD, R., C. MALLORY – SMITH, J. MICKELSON (2004): Downy brome<br />

(Bromus tectorum) response to imazamox rate and application timing in herbicide –<br />

resistant winter wheat. Weed Technology, 18, 4, 1043-1048.<br />

WOZNICA, Z., (2005): Recent advances in adjuvant formulation thechnology, 13 th EWRS<br />

Symposium, Bari – Italy, CD – ROM<br />

46


Herbologia Vol. 7, No. 1, 2006.<br />

HERBICIDE-RESISTANT CROPS - ADVANTAGES AND RISKS<br />

Zvonko Pacanoski<br />

Faculty for Agricultural Sciences and Food, 1000 Skopje, R. Macedonia<br />

zvonkop@zf.ukim.edu.mk<br />

Abstract<br />

Revealing of genetically modified, herbicide-resistant crops (HRCs)<br />

at the end of the 20 th century brought many controversies. At one side, HRCs<br />

enable the farmers to more effectively use reduced or no-tillage cultural<br />

practices, eliminate of troublesome weeds resistant to some herbicides, using<br />

nonselective herbicides, usually glyphosate and glufosinate as<br />

“environmentally friendly” herbicides, decrease of expenses for crop<br />

production, and finely, more economically and effectively manage of weeds.<br />

At the other side, there is fear of possibile developing of weed resistant to<br />

non-selective herbicides, appearance of volunteer HR crops, invasion of the<br />

environment beyond the farm boundary, influence of HRCs to biodiversity,<br />

exchange of genetic material between related HRCs and wild progenitors,<br />

conventional crops, and weeds. The coming period should clarify the<br />

eventual impact of these powerful new tools on weed science, weed<br />

management, environment and human health.<br />

Keywords: Genetically modified organisms (GMOs), herbicide-resistant crops<br />

(HRCs), glyphosate, glufosinate, weeds<br />

Introduction<br />

Weed control obtained a new dimension with recent revealing and<br />

practical application of plant growth regulators. From that moment chemical<br />

industry in collaboration with science have focused on creation of new<br />

herbicide active ingredients with large efficacy, selectivity and possibility for<br />

use in all crops against almost all weed species. In that period has developed<br />

opinion that weed problem would be dissolved forever, without investing<br />

efforts in finding and development alternative methods for weed control.<br />

However, a long period did not pass when the first problem appeared -<br />

resistance of some weeds to some herbicides, on which, until that moment the<br />

weedd had been susceptible. The number of resistant weeds to some<br />

herbicides continuously have increased. According HRAC and WSSA, about<br />

182 weed species worldwide (109 broad-leaved and 73 grass weeds)<br />

developed resistance to large number of different herbicides from all mode of<br />

action groups. In the beginning, increasing the herbicides rate was one of


Z. Pacanoski<br />

possible solution for this problem. However, the success was of short<br />

duration and partial, but harms were double. Namely, after certain period the<br />

resistance appeared again, and first environment damages concerning<br />

pollution of the soil, surface and underground water. Some widely used<br />

herbicides were out of use and forbidden, but weed problem stay remained.<br />

Science and chemical industry were faced with challenge to find alternative<br />

ways in combating weeds. Many of them (biological, physical, alellopathy)<br />

gave insufficient results, and they are not adequate substitute for herbicides.<br />

Recently, in the solving on this problem, biotechnology with genetic<br />

engineering were involved. Scientists are now creating new plants, in order<br />

to fight weeds more simpler, cheaper, and at the same time, to stop pollution<br />

of environment. All these points of view and principles are taken into<br />

consideration in creating of genetic modified organisms (GMOs)<br />

The aim of this review is to give some information about GMOs,<br />

particularly HRCs, their advantages, disadvantages and risks for plant<br />

biodiversity.<br />

From the moment creating GMOs to the time of their marketing<br />

(1995), short period passed, but their planted area year after year have been<br />

increasing. According Berca (2004), GMOs are cultivated 12% more in 2002<br />

compared to 1995, 15% more in 2003 compared to 2002, respectively.<br />

According to ISAAA, sowings of GM crops rose to 81 million hectares in<br />

2004 recording a 20 percent increase over 2003.<br />

Generally, GM crops are planted in 18 countries, at about 10 million<br />

farms. The biggest producers of GMOs (James, 2003) are: USA, Canada,<br />

Argentina, Brazil, and China. These countries account to about 98% of all<br />

world GMOs production.<br />

Tab. 1. The biggest world producers of GMO in the world (James, 2003)<br />

Country Surface mil. ha % of total<br />

cultivated GMO<br />

USA 42,8 63<br />

Argentina 13,9 21<br />

Canada 4,4 6<br />

Brasil 3,0 4<br />

China 2,8 4<br />

Tab. 2. Distribution of GMO on products (James, 2003)<br />

Crops Surface mil. ha % of total<br />

cultivated GMO<br />

Soya 41.7 61<br />

48


Herbicide-resistant crops – advatages and risk<br />

Maize 15.5 23<br />

Cotton 7.2 11<br />

Canola 3.6 5<br />

Advantages of herbicides resistant crops<br />

About 71% (more than 40 million hectares) of total cultivated GMOs<br />

in the world are herbicides resistant crops (HRCs). The reason why the<br />

biggest percent of GMOs are HRCs lie in the numerous advantages of this<br />

new technology in weed control. One of the main benefits is the possibility of<br />

controlling a range of broad-leaves and grass weeds with one or two<br />

properly timed application of glyphosate (Baldwin, 1999). Conventional<br />

herbicide method required applications of two or more different herbicides.<br />

In addition, conventional herbicide programme is tightly connected with crop<br />

and weed growth stages, also with ecological conditions (climate and soil).<br />

Taking into consideration all these factors, possibility for potentional crop<br />

injury are high (Johnson et al., 2002). Cultivating transgenic crops enable<br />

using nonselective herbicides, usually glyphosate and glufosinate, usually<br />

once, but in some cases two applications of these herbicides controlled<br />

weeds. At the same time, application period is more flexible than in<br />

conventional way; it does not depend of growth stages of crops and weeds,<br />

while the possibility for injury crop is minimal (Carpenter and Gianessi,<br />

1999).<br />

Utilization of new method of weed management reduces application<br />

of long-term residual herbicides, that are, according to new environmental<br />

standards, unacceptable (Shaw et al., 2001). New weed management<br />

technology is particularly applicable in the places with weeds dominant<br />

resistant to some selective herbicides. In these circumstances, glyphosate and<br />

glufosinate are a good solution for this problem, because efficacy of these<br />

herbicides is very high. One of these herbicides is capable, alone, to solve the<br />

weed problem, even better than standard combination of two or three<br />

selective herbicides.<br />

Numerous investigations confirmed excellent results in efficacy of<br />

new way of weed control, which is usually more than 95%. Since neither<br />

glyphosate nor glufosinate have residual activity after their application,<br />

second application, if necessary, is possible for achieving weed-free crops.<br />

The other major advantage of using glyphosate and glufosinate in<br />

HRCs is their role in no-till and zero-till agriculture. Minimizing or<br />

eliminating soil tillage reduces or prevents soil erosion, humus degradation<br />

and destruction of soil structure. However, the weeds soon become main<br />

problem in such a system if previous mentioned herbicides are not used.<br />

49


Z. Pacanoski<br />

In context of this thesis are farmer’s reactions cultivating transgenic<br />

maize and soybean resistant to glyphosate and glufosinate under no-till<br />

conditions. Their reactions are mainly positive and they confirm that using<br />

transgenic crops give bigger opportunities in better weed management system<br />

with lower herbicide inputs.<br />

Beside excellent efficacy in weed control, glyphosate is<br />

environmentally friendly herbicide. After glyphosate is applied to forests,<br />

fields, and other land by spraying, it is strongly adsorbed to soil, remains in<br />

the upper soil layers, and has a low propensity for leaching. Glyphosate<br />

readily and completely biodegrades in soil. Its average half-life in soil is<br />

about 60 days; main product of its degradation is aminomethylphosphoric<br />

acid, which is broken down further by soil microorganisms as well.<br />

Glyphosate is practically non-toxic for birds, mammals and bees (LD 50 for<br />

rats is 5600 mg/kg. Oral LD 50 for rabbits and goats is more than 10,000<br />

mg/kg)<br />

Similar is the situation for glufosinate, as nonselective contact and<br />

nonpersistent herbicide with moderate leaching. Glufosinate is also<br />

environmentally friendly and practically non-toxic (oral LD 50 above 2000<br />

mg/kg and dermal 4000 mg/kg). It has no residual activities and does not<br />

inflict restriction to crop rotation. Half-life is about 40 days.<br />

Toxicity<br />

category<br />

Tab. 3. Ecotoxicological categories<br />

Mammalian<br />

(acute oral)*<br />

mg/kg<br />

Avian<br />

(acute oral)*<br />

mg/kg<br />

Avian<br />

(dietary) _<br />

ppm<br />

Aquatic<br />

organisms ‡<br />

ppm<br />

Very highly toxic 2000 >5000 >100<br />

Except for weed control, glyphosate and glufosinate are efficient<br />

against some plant pathogens. For instance, glufosinate inhibits infection of<br />

glufosinate-resistant creeping bentgrass (Agrostis palustris) with several<br />

plant pathogens (Liu et al., 1998). Also, HRCs can be especially useful for<br />

eradication of parasitic weeds (Joel et al. 1995). According Altman (1993,<br />

cited by Duke, 1999), more additional and profound investigations need to be<br />

done on secondary effects of these herbicides in order to fully determine their<br />

roles in integrated pest management.<br />

Taking into consideration all previous mentioned advantages of new<br />

technology, herbicide industry appears to be rapidly transforming from a<br />

50


Herbicide-resistant crops – advatages and risk<br />

chemically based to biotechnology oriented industry. The largest pesticide<br />

producers of the US and Europe have invested heavily in plant biotechnology<br />

and seed industry. Every year, since the first experimental releases in 1987,<br />

HRCs have accounted for nearly one-third of field tests conducted under<br />

USA authority. The biggest part of trials are conducted by companies experts<br />

who have created and applied new technologies. Successful creation of new<br />

HRCs will be economical method of expanding the market for products for<br />

which companies have already have sophisticated equipment and highly<br />

specialized staff.<br />

Possible risks and concern of GMOs<br />

As every new technology, also GMOs, beside positive, may have<br />

some potentially negative aspects, which are, fortunately, still in the domain<br />

of speculation. Controversies surrounding GMOs commonly focus on human<br />

and environmental safety, labelling and consumer choice, intellectual<br />

property rights, ethics, food security, poverty reduction and environment<br />

conservation. Some of them are real and the experts are aware for that. They<br />

take every measure to eliminate them on time. There is not any reason for<br />

concern. The following text will elaborate some of the potential risks and<br />

negative implication from HRC technology.<br />

One of the real risk in cultivation of HRCs is possibility for<br />

development of weed resistant mechanism to non-selective herbicides.<br />

Namely, application of unilateral strategy in weed control management and<br />

multiple glyphosate and gluphosinate application presents serious threat in<br />

the process for appearing of resistant weeds (Derksen et al., 1999). Weed<br />

resistance to glyphosate and gluphosinate, according some authors<br />

(Bradshow et al., 1997; Waters, 1991) is less probable, and will be rarer<br />

evolved to glyphosate and gluphosinate than to many other herbicides. The<br />

reason for this, according to these authors, is slow development of functional<br />

genes of resistance and the rich biodiversity in agroecosystem where these<br />

herbicides are applied. Nevertheless, glyphosate-resistant weeds have<br />

appeared, first in rigid ryegrass (Lolium rigidum Gaudin). in Australia<br />

(Powels et al., 1998; Pratley et al., 1999), then goosegrass (Eleusine indica L.<br />

Gaertn) in Malaysia, and subsequently, hairy fleabane (Conyza banariensis<br />

LO. Cronq.) and horseweed (Conyza canadensis L. Cronq.), in the USA<br />

(Heap, 2001). Resistance to glyphosate in Australian ecosystem is result of<br />

long-term application of this herbicide in no-till system. No-till system has<br />

poor diversity than conventional systems. Weed resistance to glyphosate in<br />

the USA is recorded in HR soybean because of continuous use in so-called<br />

minimal diversity system (Powels, 2003). It is clear that weed resistance to<br />

glyphosate in previous mentioned agroecosystems is due to poor diversity<br />

51


Z. Pacanoski<br />

and unilateral weed control management. More scientists who study this<br />

weed-resistant phenomenon, agree about possibility for escalation of this<br />

problem and they emphasize importance for designing appropriate system to<br />

prevent development of weed-resistance in HRC system. Increased diversity<br />

i.e., a wider range of rotational crops, diversity of herbicides used (herbicides<br />

with different mode of action, for example sulfentrazol + glyphosate in<br />

soybean) (Krausz et al., 2003), better cropping practices leading to more<br />

competitive crops, and use of nonherbicide weed control, can reduce the<br />

dependence on glyphosate and reduce the likelihood of resistance developing.<br />

Other risk for this system are “volunteer HR crops”. Long-term use of<br />

HRCs, particularly in crop rotation system only with HRCs, can be a serious<br />

problem. Namely, the volunteer plants of previous HR crop in the next HR<br />

crop can be problem, if the next HR crop is resistant to the same herbicide<br />

like the previous one. To avoid this real risk, it should be applied a preventive<br />

strategy such as presowing/preplanting soil cultivation and application of<br />

alternative soil herbicides. The herbicides choice should be done very careful,<br />

because of possibility of multiple and crossing herbicide resistance. The best<br />

way to prevent this undesirable appearance is application of herbicide with<br />

different mode of action than glyphosate and glufosinate. Appearance of<br />

volunteer HR plants is concerned particularly in harvest of transgenic rice<br />

and soybean, when some seeds can shed on the soil during this operation.<br />

Introduction of new technology impose the question for possible<br />

influence of HRCs to biodiversity. It can not give precise answer on this<br />

question, although numerous investigations with HRCs are made all over the<br />

world. Maybe we can answer this question indirectly, toward conventional<br />

crop production and its influence to biodiversity.<br />

The most important crop grown in Brazil is soybean, with<br />

approximately 13 million ha planted. The most important weeds associated<br />

with the crop belong to families Poaceae, Amaranthaceae, Cyperaceae,<br />

Euphorbiaceae and Asteraceae (Foloni and Christoffoleti, 1999) (cit. by<br />

Riches and Valverde, 2002). Agriculture intensification, adoption of nontillage<br />

systems, and overreliance on herbicides has resulted in the increased<br />

occurrence of hard to kill broad-leaf weeds, including those in the genera<br />

Cammelina, Euphorbia, Ipomoea, Borrerlia and Tridax (Merotto et al.,<br />

1999). Additionally, wild poinsettia (Euphorbia heterophylla L.) and hairy<br />

beggarticks (Bidens pilosa L.) have evolved resistance to ALS inhibitors<br />

(Ponchio et al., 1997; Theisen et al., 1997; Vidal et al. 1997). In Argentina,<br />

pigweed (Amaranthus quitensis H.B.K.) evolved resistance to imazethapyr<br />

and chlorimuron in soybean (Christoffoleti et al.,1997, cited by Riches and<br />

Valverde, 2002). Non-tillage systems and unilateral ALS-herbicide<br />

application changed biodiversity in the soybean.<br />

52


Herbicide-resistant crops – advatages and risk<br />

According to the fact that weeds are an important component of<br />

agrobiodiversity, shifts in the composition of weed flora provoke change of<br />

biodiversity (Spahillari et al.,1999). The impact of HRCs to biodiversity will<br />

be the same. According to Forcella (1999), it could be anticipated that most<br />

HRCs are associated with nonresidual herbicides and the promotion of zerotillage,<br />

the seed bank density of weeds with protracted emergence periods or<br />

of those emerging late in the growing season is more likely to increase as<br />

compared with that of other species more easily controlled.<br />

Changes in the type and frequency of herbicide application in HRCs<br />

will provoke gradual removal on perennial and colonization of annual weeds<br />

(Sweet et al., 1999). For biodiversity conservation and avoiding undesirable<br />

effects, choice of complex measures of weed control (integral weed<br />

management) is needed, particularly regular and timely soil cultivation, crop<br />

rotation and application of herbicides with different mode of action.<br />

One of the potential questions is: can HRCs become invasive beyond<br />

the field boundary Numerous experts agree that HRCs can not survive out of<br />

the agroecosystem’s boundary. This thesis is corroborated with the fact that<br />

in the process in their selection with application of molecular biotechnology,<br />

and then in the process of domestication, HRCs became completely depended<br />

on human activities. For example, transgenic varieties of soybean and maize<br />

which are cultivated in South America, according to Colwell (1994), are<br />

weak competitors to plants species out of the arable land.<br />

However, the possibility for exchange of genetic material between<br />

related HRCs and wild ancestors, conventional crops and weeds, concern<br />

scientists most. This phenomenon is very possible in the centers of their<br />

origin. Geographic distance and pollination system are key factors in existing<br />

of probability for genes exchange. Soybean, the first and the most cultivated<br />

worldwide HRC, is exotic plant, whose wild progenitors originated from<br />

China, Russia, Japan, and Korea (Palmer et al., 1996). That means, if HR<br />

soybean is cultivated out of its origin centres there is not threat for gene<br />

exchange and, in the same time, for plant biodiversity.<br />

Existing and real risk is possibility of exchange of genetic material<br />

between transgenic and conventional crops. Particularly this problem is<br />

stressed for farms certified for so-called organic production. In order to avoid<br />

pollen transmission from transgenics crops to crops for organic, and also, for<br />

conventional production, space isolation is necessary. But, still there does not<br />

exist precise regulative, which will define minimum distance between<br />

trangenics and other crops.<br />

The most serious threat for new technology is possibility for gene<br />

exchange between related HRCs and weeds. Many different sources have<br />

been consulted and analyzed in order to estimate risk of appearance of<br />

hybrids between HRCs and weeds. According to scientific investigations of<br />

53


Z. Pacanoski<br />

Keeler et al. (1996) and Arriola (2000) based on numerous research activities<br />

in Europe and North America, it is clear that HR crops and related weedy<br />

plants can exchange genes through pollen transfer. The identification of<br />

spontaneous hybrid forms in a number of crop-weed complexes is well<br />

established, including between Sorghum halepense (L.) Pers. and Sorghum<br />

vulgare Pers., and between wild and cultivated forms of Helianthus annus<br />

L.and Oriza sativa L. (Arias and Reisberg, 1994; Arriola and Ellstrand,<br />

1996). According to Dale (1994), possibility for gene exchange between<br />

HRCs and weed population depends of three factors:<br />

i) sexual compatibility between crop and weed<br />

ii) possibility of spontaneous exchange of genetic material<br />

iii)<br />

between crop and weed (spontaneous hybridization)<br />

the manner in which new characteristic in the crop-weed<br />

hybrid will behave under environmental conditions<br />

Formation of hybrids between HRCs and weeds will provoke<br />

difficulties in weed control, similar to those with herbicides resistant<br />

weeds in conventional crop production. The sexual transfer of genes<br />

from crops to weeds, as it was mentioned before, is probably the<br />

biggest risk for the environment and can limit the cultivation of<br />

transgenic species. It is case of weeds related to transgenic canola<br />

which transfers her pollen towards wild species of Brassica, Sinapis<br />

etc. To avoid this phenomenon, today a specific management is<br />

practiced, which involve:<br />

i) maternal inheritance<br />

ii) male sterility<br />

iii) seed sterility<br />

iv) cleistogamy<br />

v) apomixis<br />

vi) incompatible genomes<br />

If maternal inheritances are presented, the interest gene is expressed<br />

only in chloroplasts and can not be dispersed through pollen to nontransgenic<br />

plants. Male sterility (sterility of pollen) is very important to<br />

eliminate the crossings in the environment (outcrossing). Therefore,<br />

transgenic plants should be created by getting seeds where the interest gene<br />

should be dominated in the mother plants, and pollination should be done<br />

with a non-transgenic line.<br />

At cleistogamic plants pollination is made inside the flowers before<br />

these open. In this way the external crossing can be avoided (the case of<br />

Triticum durum).<br />

Apomixis is an asexual type of reproduction in which the plant<br />

embryos grow from egg cells without being fertilized by pollen—the male<br />

54


Herbicide-resistant crops – advatages and risk<br />

part of the plant. The result is production of seed which inherited only<br />

maternal genes, and the plants are identical with maternal plant.<br />

Conclusions<br />

Taking into consideration all advantages, disadvantages and risks<br />

which are connected with the new technology, particularly with HRCs,<br />

undoubtly one question is posed: how and to what extent is possible use<br />

HRCs without causing undesirable effects In order to be controlled, the new<br />

technology should be regulated by low. That means preparation and<br />

verification of international GMOs regulations and norms, possibilities of the<br />

pesticide-biotechnology industry to protect and recoup their investments<br />

Several countries have already established legislation on the release<br />

of HRCs which are product of the new technology. The EU states in the<br />

recently revised Directive 2001/18/EEC states that GMOs have to undergo a<br />

scientific assessment of risks to human health and environment. The<br />

Directive presents a framework for the risk assessment, which appears to be<br />

derived from the framework put forward by the joint consultation on food<br />

safety brought up by the World Health Organization (WHO) and Food and<br />

Agriculture Organization (FAO).<br />

It is clear, that HRCs are (or soon will be) strongly impacting weed<br />

management choice. Their mass use will decrease the cost of effective weed<br />

management in the short or medium term. Their use will speed up adoption<br />

of reduced and no-tillage agriculture, greatly reducing the environmental<br />

damage by farming by reducing soil erosion (wind and water) and by<br />

reducing the use of herbicides likely to be found in surface and ground water.<br />

Herbicides resistance and new weed species that arise as a result of this<br />

technology will be dealt with traditional methods, such as rotating and<br />

mixing herbicides and rotating crops.<br />

References<br />

ARIAS, D.M., RIESBERG, L.H. (1994): Gene flow between cultivated and wild sunflowers,<br />

Theory of Applied Genetic, 89, 655-660<br />

ARRIOLA, P.E. (2000): Risks of escape and spread of engineered genes from transgenic<br />

crops to wild relatives, Biosafety Reviews, Biosafety Information Network and<br />

Advisory Service, http:/www.binas.unido.org.binas/library<br />

ARRIOLA, P.E., ELLSTRAND, N.C. (1996): Crop-to-weed gene flow in the genus<br />

Sorghum (Poaceae): spontaneus interspecific hybridisation between johnsongrass,<br />

Sorghum halepense, and crop sorghum, S. bicolor, American Journal of Botany, 83,<br />

1, 153-160<br />

BALDWIN, F.L. (1999): The value and exploatation of herbicide-tolerant crops in the US.<br />

Proc.Br.Crop Prot. Conf.Weeds 635-660<br />

55


Z. Pacanoski<br />

BERCA, M. (2004): Perspectives Regarding Weeds Control, Universirty Foundation CERA<br />

for Agriculture and Rural Development<br />

BRADSHOW, L.D., PADGETE, S.R., KIMBALL, S.L., WELLES, B.H. (1997):<br />

Perspectives on glyphosate resistance, Weed Technology, 11, 189-198<br />

CARPENTER, J., GIANESSI, L. (1999): Herbicide-tolerant soybeans: why growers are<br />

adopting Roundap ready varietes, AgroBioForum, 2, 65-72<br />

CLAYTON, G.W., NEIL-HARKER, K., O’DONOVAN, J.T., BAIG, M.N., KIDNIE, M..J.<br />

(2002): Glyphosate timing and tillage system effects on gliphosate-resistant canola<br />

(Brassica napus), Weed Technology, 16, 124-130<br />

COLWELL, R.K. (1994): Potential ecological and evolutionary problems of introducing<br />

transgenic crops into the environment, Sharing Biotechnology Regulatory<br />

Experiences of the Western Hemisphere, Ithaca, NY: International Service for the<br />

Acquisition for Agri-Biotech Applications, 33-46<br />

DALE, P.J. (1994): The impact of hybrids between geneticaly modified crop plants and their<br />

related species: general considerations, Molecular Ecology, 3, 31-36<br />

DERKSEN, D.A., HARKER, K.N., BLACKSHOW, R.E. (1999): Herbicide-tolerant crops<br />

and weed population dynamics in western Canada, Proceedings Brighton Crop<br />

Conference-weeds, Farnham, Surrey, U.K. British Crop Protection Coucil, pp. 417-<br />

424<br />

DUKE, O. S. (1999): Weed Management: Implications of Herbicides Resistant Crops,<br />

USDA-ARS-Natural Products Utilization Research Unit,<br />

http://www.nbiap.vt.edu/proceedings.duke.html<br />

FOLONI, L.L., CHRISTOFFOLETI, P.J. (1999): Chemical weed control in soybeans in<br />

Brazil using new herbicides and mixtures, Proceedings Brighton Crop Conferenceweeds<br />

315-318<br />

FORCELLA, F. (1999): Weed seed bank dynamics under herbicides tolerant crops,<br />

Proceedings Brighton Crop Conference-weeds 409-416<br />

HEAP, I..M. (1997): The occurence of herbicide-resistant weeds worldwide, Pesticides<br />

Science, 51, 235-243<br />

JAMES, C. (2003): Preview global review of commercialized transgenetic crops, Ithaca, NY,<br />

ISAAA, briefs, No. 21, Preview ISAAA, 17 p.<br />

JOEL, D.M., KLEIFELD, Y., LOSNER-GOSHEN, D., HERZLINGER, G., GRESSEL, J.<br />

(1995): Transgenic crops to fight parasitic weeds, Nature, 374<br />

KEELER, K.H., TURNER, C.E., BOLICK, M.R. (1996): Movement of crop trasgenes into<br />

wild plants, Agricultural, Environmental, Economic, Regulatory and Techical<br />

aspects, Boca Raton, FL:CRC Lewis, 303-330<br />

KRAUSZ, R.F., YOUNG, B.G. (2003): Sufentrazone enhances weed control of glyphosate<br />

in glyphosate-resistant soybean (Glycine max.), Weed Technology, 17, 249-255<br />

LIU, C.A., ZHONG, H., VARGAS, J.,PENNER, D., STICKLEN, M.(1998): Prevention of<br />

fungal dieseases in transgenic, bialaphos and glfosinate resistant creeping bentgrass<br />

(Agrostis palustris), Weed Science, 46, 139-146<br />

MADSEN, K.H., JENSEN, J.E. (1995): Weed control in glyphosate-tolerant sugarbeet (Beta<br />

vulgaris), Weed Research, 35, 105-111<br />

MADSEN, K.H., VALVERDE, B.E., JENSEN, J.E. (2002): Risk assessment of herbicideresistant<br />

crops: A Latin America perspective using rice (Oryza sativa) as a model,<br />

Weed Technology, 16,215-223<br />

MEROTTO, A., VIDAL, R.A., FLECK, N.G. (1999): Soybean tolerance to syntetic auxin<br />

and potential of mixures with protox-inhibiting herbicides, Proceedings Brighton<br />

Crop Conference-weeds 319-324<br />

56


Herbicide-resistant crops – advatages and risk<br />

PALMER, R.G., HYMOVITZ, T., NELSON, R.I. (1996): Germplasm diversity wirthin<br />

soybean, Genetics, Molecular biology and Biotechnology, Wallingford, UK: CABI<br />

Publishing , 1-35<br />

POWELS, S.B., LORRAINE-COLWILL, D.F., DELLOW, J.J.,PRESTON, C. (1998):<br />

Evolved resistance to glyphosate in rigid ryegrass (Lolium rigidum) in Australia,<br />

Weed Science, 46, 604-607<br />

RICHES, C.R., VALVERDE, B.E. (2002):Agricultural and Biological Diversity in Latin<br />

America: Implications for Development, Testing and Commercialization of<br />

Herbicide-Resistant Crops, Weed Technology, 16, 200-214<br />

SHOW, D., ARNOLD, J.C., SNIPES, C.E., LAUGHLIN, D.H., MILLS, J.A. (2001):<br />

Comparison of glyphosate-resistant and nontransgenic soybean (Glycine max.)<br />

herbicides system, Weed Technology, 15, 676-685<br />

SHAW, D.R., ARNOLD, J.C. (2002): Weed control from herbicide combination with<br />

glyphosate, Weed Technology, 16, 200-214<br />

SHAW, D.R., BRAY, C.S. (2003): Foreign material and seed moisture in glyphosateresistant<br />

and conventional soybean system, Weed Technology, 17,389-393<br />

SPAHILLARI, M., HAMMER, K., GLADIS, T.,DIEDERICHSEN, A. (1999): Weeds as<br />

part of agrobiodiversity, Outlook Agriculture, 28, 227-232<br />

SCOTT, G.H., ASKEW, S.D., WILCUT. J.(2002): Glyphosate systems for weed control in<br />

glyphosate-tolerant cotton (Gossypium hirsutum), Weed Technology, 16, 191-198<br />

SQUIRE, G.R., CRAWFORD, J.W., RAMSAY, G., THOMPSON, C., BROWN, J. (1999):<br />

Gene flow at the landscape level; Relevance for transgenetic crops, British Crop<br />

Protection Council, pp. 241-246<br />

SWEET, J.B., NORRIS, C.E., FLECK, N.G. (1999): Assessing the impact and consequences<br />

of the release and commercialisation of GMCs, British Crop Protection Council,<br />

241-246<br />

57


Herbologia Vol. 7, No. 1, 2006.<br />

PRODUCTION OF ALLERGENIC POLLEN BY RAGWEED (AMBROSIA<br />

ARTEMISIIFOLIA L.) IS INCREASED IN CO 2 -ENRICHED<br />

ATMOSPHERES<br />

Taib Šarić 1 , Ivica ðalović 2<br />

1 Faculty of Agriculture, Sarajevo, Bosnia&Herzegovina, e–mail: tsaric@bih.net.ba<br />

2 Faculty of Agriculture, Čačak, Serbia&Montenegro<br />

Abstract<br />

Ragweed (Ambrosia artemisiifola L.) is tipically a pioneer plant,<br />

which invades recently disturbed soils. It belongs to the group of the annual<br />

plants that emit pollen during late summer and early fall near the end of the<br />

growing season. This pollen is produced in such large quantities that is a key<br />

contributor to ex-Yugoslavia hay–fever problems near the end of summer.<br />

Anemophilous pollen grains are airborne and meteorological factors<br />

have great impact on their dispersal. Precipitation, wind speed, temperature<br />

and air moisture are often cited as influencing airborne pollen concentrations<br />

(Emberin et al., 1996). Numerous studies have already dealt with these<br />

aspects. Recently, Bartkova et al. (2003) showed that temperature has a<br />

marked impact on ragweed pollen production in Bratislava, Slovakia.<br />

Relative humidity is an important factor as well but not quite as determining.<br />

In the same direction, Barnes et al. (2001) noticed that usual weather<br />

conditions, temperature and relative humidity have onl little influence on the<br />

day–to–day variation of ragweed pollen counts. However, unstable<br />

atmospheric conditions such as the crossing of a cold front has the greatest<br />

impact of all the weather–related events on airborne ragweed pollen counts.<br />

According to the authors, only heavy rainfall has a distinct impact on pollen<br />

concentrations. Peak pollen production has been described to occur shortly<br />

after sunrise and may be related to photo cycle periods or cooler morning<br />

temperatures and lower humidity. Different statistically based types of<br />

models trend to predict pollen concentrations from meteorological conditions<br />

as reported by Laiidi et al. (2003). Bringfelt (1982) subsequently published<br />

some correlation studies dealing with pollen concentrations and weather<br />

parameters for forecasting purposes. One of his main conclusions is that daily<br />

temperature values from early spring have a major influence on the timing of<br />

the beginning of the pollen season and subsequently on its day–to–day<br />

variation.<br />

Keywords: ragweed, aeroallergens, allergy, pollen, CO 2, climate change


T. Šarić and I. ðalović<br />

Introduction<br />

Recent studies have shown a link between warming trends within the<br />

past 50 years and the phenology and abundance of allergenic pollen released<br />

by a number of European tree species (Jaeger et al., 1996; Emberlin et al.,<br />

1997). However, only limited data are currently available to evaluate the<br />

direct effects of rising atmospheric CO 2 concentrations on pollen production<br />

by allergenic plants and its potential impact on public health (Ziska and<br />

Caulfield, 2000).<br />

Human allergic responses to the pollen of certain plant species (hay<br />

fever and allergenic rhinitis) is a serious environmental health issue (National<br />

Institute of Health, 1993). Aeroallergens, including pollen, also play a role in<br />

the exacerbation of asthma (D'Amato et al., 1994). The prevalence of both<br />

hay fever and asthma has increased significantly in recent decades (Wuthrich,<br />

1991; Arrighi, 1995). Little research has been devoted to understanding how<br />

various components of global environmental change influence allergenic<br />

pollen production and, thus, the potential for pollen-related disease.<br />

An increase in the concentration of atmospheric CO 2 is one of the<br />

most certain predictions of climate change models. CO 2 concentration has<br />

increased by 29% since preindustrial times, and is expected to double again<br />

sometime between 2050 and 2100 (Houghton, 1996).<br />

Plants grown in CO 2 -enriched atmospheres generally grow faster and<br />

are larger at maturity, although the magnitudes of growth and physiological<br />

enhancements vary considerably with environmental conditions and species<br />

identity (Bazzaz, 1990; Curtis and Wang, 1998). In one recent study, Ziska<br />

and Caulfield (2000) found that exposing ragweed plants to the higher CO 2<br />

concentrations predicted in the year 2100 doubled the quantity of pollen<br />

produced.<br />

Ragweed is a plant common to roadsides and disturbed habitats<br />

throughout most of the United States and Canada (Basset and Crompton,<br />

1975). It has male and female flowers born on distinct axillary branches,<br />

allowing for independent control of allocation to sexes (Payne, 1963).<br />

Throughout its distribution, ragweed pollen is one of the most abundant<br />

aeroallergens in late summer and fall, and it is one of the primary causes of<br />

seasonal pollen allergy in North America (Lewis et al., 1983).<br />

Consequently, ragweed pollen and specific allergens extracted from it<br />

have been used in many clinical studies, and the biochemistry and genetics of<br />

ragweed allergens and their impacts on the human immune systems are well<br />

understood (Griffith et al., 1991; Naclerio et al., 1997).<br />

This study investigates the direct impact of rising CO 2 concentrations<br />

on pollen production of ragweed. The results can be used to more accurately<br />

60


Production of allergenic pollen by ragweed (Ambrosia artemisiifolia L.) is increased in ...<br />

evaluate the future risks of hay fever and respiratory disease exacerbated by<br />

allergenic pollen, and to develop strategies to mitigate them.<br />

Pollen<br />

Most pollen grains are smaller than 80 microns which is about the<br />

width of a human hair. Some pollens cause allergic reactions in some people.<br />

Pollen allergies are specific to the individual and the pollen type. Pollen<br />

grains that are wind-dispersed cause most of the allergy problems because<br />

these are the pollens that make their way into our nasal passages and eyes.<br />

Pollens from some plants, such as roses and tulips, are too big to be<br />

transported by the air alone. They depend on insects and birds to carry the<br />

pollen from plant to plant.<br />

Most pollens are spherical in shape. Ragweed pollens, which bothers<br />

many people, appear to have spikes or other similar features.<br />

While grass pollen is most prominent in May and continues into<br />

September, most weed pollen quantities increase throughout May and into<br />

June, ragweed pollen is very significant from late August through mid-<br />

September. The first hard frost of autumn typically brings our pollen season<br />

to a close in October. Pollen seasons can last for several months.<br />

On average, grasses are the most potent pollen grain on human<br />

allergies. To put this in perspective, here is an example. If the air had a<br />

pollen density of 90 grains/m 3 for trees, it would take a density of only 20<br />

grains/m 3 for grasses to cause the same level of symptoms associated with<br />

allergies. It would take 50 grains/m 3 for weeds. On a grain for grain basis,<br />

grass pollen has a higher allergic potential than either tree or weed pollen.<br />

Because of these differences in allergic potential, it takes a higher density of<br />

weed or tree pollen to constitute a "high" pollen rating than it takes for grass<br />

pollen.<br />

Some studies have examined trends in pollen amount over the latter<br />

decades of the 1900s and found increases to be associated with local rises in<br />

temperature (Corden and Millington, 2001; Spieksma et al., 1995). Changes<br />

in climate appear to have altered the temporal and spatial distribution of<br />

pollen. For example, some studies have found that trends toward earlier<br />

pollen seasons are associated with local warming over the latter decades of<br />

the 1900s (Emberlin et al., 2003; Fitter and Fitter 2002), and recent reports<br />

have concluded that the duration of the pollen season is extended in some<br />

species (Huynen and Menne 2003). Finally, several studies have examined<br />

other attributes of allergenic plants, which have also been responsive to CO 2<br />

concentration and/or temperature increases (e.g. Menzel 2000; Wulff and<br />

Alexander, 1985). These latter studies provide indirect evidence of impacts of<br />

climate change on pollen aeroallergens.<br />

61


T. Šarić and I. ðalović<br />

Impacts of climate change on aeroallergens: past and future<br />

A number of studies have revealed potential impacts of climate<br />

change on aeroallergens that may have enormous clinical and public health<br />

significance.<br />

Human activities have resulted in increases in the concentrations of<br />

atmospheric greenhouse gases and changes in global climate. Before the<br />

Industrial Era (circa 1750) atmospheric carbon dioxide (CO 2 ) concentration<br />

was 280±10 ppm for several thousand years (Prentice et al., 2001). It has<br />

risen since then to 373 ppm (Keeling and Whorf, 2004).<br />

It is estimated that global average surface temperature changed<br />

increased by 0.6 0 C since the late 19th century. Projected CO 2 concentration<br />

by 2100 ranged from 541 to 970 ppm (approximately 1.9 and 3.5 times the<br />

pre-industrial concentration). Global average surface temperature is projected<br />

to rise over the period 1990–2100 under all scenarios, ranging from 1.4 0 C to<br />

5.8 0 C (Cubasch et al., 2001).<br />

Climate change is likely to have impacts on hayfever (allergic<br />

rhinitis) and asthma via its impacts on pollens and other aeroallergens.<br />

Atmospheric variables that may have impacts on these allergens<br />

include CO 2 concentration, temperature, rainfall, humidity, and wind speed<br />

and direction. With allergic diseases already being a significant public health<br />

issue in many countries, the potential for any adverse impact resulting from<br />

climate change is of serious concern.<br />

Pollen amount<br />

Impacts of climate change on aeroallergens<br />

A number of studies have found increases in pollen associated with<br />

increases in CO 2 concentration and/or temperature. Ziska and Caulfield<br />

(2000) found pollen production of common ragweed increased significantly<br />

both from pre-industrial to current and current to the future CO 2<br />

concentration. Similarly, Wayne et al. (2002) found a significant increase in<br />

ragweed pollen production under an approximate doubling of the<br />

atmospheric CO 2 concentration, although the increase was smaller than that<br />

found previously by Ziska and Caulfield (2000), who had examined a smaller<br />

increase in CO 2 concentration from current to future.<br />

The impact of climate change on pollen production of common<br />

ragweed (Ambrosia artemisiifolia) has been assessed by Ziska et al. (2003),<br />

who used an existing CO 2 /temperature gradient between rural and urban<br />

areas. The higher CO 2 concentration and air temperature of the urban area<br />

resulted in ragweed that produced significantly greater pollen than that at the<br />

rural areas. Ragweed also flowered earlier in urban than in the rural areas.<br />

62


Production of allergenic pollen by ragweed (Ambrosia artemisiifolia L.) is increased in ...<br />

Significantly stronger allergenicity was found in the pollen from<br />

plants grown at the higher temperature.<br />

Plant and pollen distribution<br />

The potential for changes in the distribution of allergen producing<br />

species has been recognized since the early days of climate change and<br />

human health work (Last and Guidotti, 1991). It has been suggested that<br />

areas of perennial ragweeds are likely to extend, and northern colonies of<br />

annual ragweed, such as that in the UK, will probably become more<br />

persistent (Emberlin, 1994).<br />

There is some evidence that vegetational response to abrupt climate<br />

change, such as that expected over the coming decades, may be rapid (Peteet,<br />

2000). Weber and Mother (2002) recently suggested that one of the<br />

implications of increased pollen production associated with increased CO 2<br />

concentration could be more efficient wind pollination and, ultimately,<br />

greater propagation of the plant species.<br />

Research challenges<br />

The research done to date is of concern for at least two reasons. First,<br />

it suggests that the future aeroallergen characteristics of our environment may<br />

change considerably as a result of climate change, with the potential for more<br />

pollen (and mould spores), more allergenic pollen, an earlier start to the<br />

pollen (and mould spore) season, and changes in pollen distribution. Second,<br />

it demonstrates climate change has probably already had impacts on<br />

aeroallergens. However, further work is required. Study of the impacts of<br />

climate change on aeroallergens and related diseases presents many<br />

challenges. Some of these challenges, along with suggestions for further<br />

work, are outlined in this section.<br />

Land-use changes will be a significant factor in determining future<br />

aeroallergen, particularly pollen characteristics. For example, Emberlin<br />

(1994) has suggested that decreases in grassland and cereals in Europe would<br />

lead to decreases in grass pollen, and that the projected increase in oil crops<br />

such as oil seed rape (Brassica species) may lead to a greater aeroallergen<br />

load.<br />

Although further work is required in this area, with the evidence to<br />

date, it would seem prudent to consider alternative adaptive strategies. One<br />

adaptive strategy would be tighter management of a number of the allergenic<br />

plant species discussed in this article. For example, government authorities<br />

could consider more carefully which plant species are used in populated<br />

areas. It is important that public health authorities and allergy practitioners be<br />

63


T. Šarić and I. ðalović<br />

aware of these changes in the environment, and that research scientists<br />

embrace the challenges that face further work in this area.<br />

References<br />

ARRIGHI, H. M. (1995): US asthma mortality: 1941 to 1989. Ann Allergy Asthma<br />

Immunol;74:321-326. Medline.<br />

BARNES, C., PACHECO, F., LANDUYT, J., HU, F. AND PORTNOY, J. (2001): The<br />

effect of temperature, relative humidity and rainfall on airborne ragweed pollen<br />

concentrations. Aerobiologia, 17: 61–68.<br />

BARTKOVA–SCEVKOVA, J. (2003): The influence of temperature, relative humidity and<br />

rainfall on the occurrence of pollen allergens (Betula, Poaceae, Ambrosia<br />

artemisiifolia) in the atmosphere of Bratislava. (Slovakia). International Journal of<br />

Biometeorology, Published online: 11/04/2003.<br />

BASSET I. J, CROMPTON C. W. (1975): The biology of Canadian weeds. 11. Ambrosia<br />

artemisiifolia L. and A. psilostachya D. C. Can. J. Plant. Sci. 55:463-476.<br />

BAZZAZ F. A. (1990): The response of natural ecosystems to the rising global CO 2 levels.<br />

Annu Rev. Ecol. Syst. 21:167-196.<br />

BEGGS P. J. (2004): Impacts of climate change on aeroallergens: past and future. Clin Exp<br />

Allergy 34:1507–1513.<br />

BRINGFELT (1982): Plant respiratory responses to elevated CO 2 partial pressure. In: Allen<br />

L. H, Kirkham M. B, Olszyk D. M, Whitman C, editors. Advances in CO 2 Effects<br />

Research. Madison, WI: American Society of Agronomy, 1-29.<br />

CORDEN J. M, MILLINGTON W. M. (2001): The long-term trends and seasonal variation<br />

of the aeroallergen Alternaria in Derby, UK. Aerobiologia 2001; 17:127–36.<br />

CUBASCH, U, MEEHL, G. A, BOER G. J. (2001): Projections of future climate change. In:<br />

Houghton J. T, Ding Y, Griggs D. J, Noguer M, van der Linden PJ, Dai X, Maskell K,<br />

Johnson CA, eds. Climate change 2001: the scientific basis, Contribution of Working<br />

Group I to the Third Assessment Report of the Intergovernmental Panel on Climate<br />

Change. Cambridge: Cambridge University Press, 2001; 525–82.<br />

CURTIS PS, WANG X. (1998): A meta-analysis of elevated CO2 effects on woody plant<br />

mass, form, and physiology. Oecologia, 113:299-313.<br />

D'AMATO G, GENTILI M, RUSSO M. (1994): Detection of Parietaria judaica airborne<br />

allergenic activity: comparison between immunochemical and morphological<br />

methods including clinical evaluation. Clin Exp Allergy 24:566-574. Medline.<br />

EMBERLIN J, DETANDT M, GEHRIG R, JAEGER S, NOLARD N, RANTIO-<br />

LEHTIMAÄKI A. (2003): Responses in the start of Betula (birch) pollen seasons to<br />

recent changes in spring temperatures across Europe. Int. J Biometeorol. 2002;<br />

46:159–70.<br />

EMBERLIN J, MULLINS J, CORDEN J. (1997): The trend to earlier birch pollen seasons in<br />

the U.K.: abiotic response to changes in weather conditions Grana; 36:29-33.<br />

EMBERLIN J. (1994): The effects of patterns in climate and pollen abundance on allergy.<br />

Allergy; 49:15–20.<br />

EMBERLIN, J. AND NORRIS–HILL, J. (1996): The influence of wind speed on the<br />

ambient concentrations of pollen from Gramineae, Plantus and Betula in the air of<br />

London, England. In: M. Muilenberg and H. Burge (eds), Aerobiology, CRC Press,<br />

Lewis Publishers, New York, pp. 27–38.<br />

FITTER A. H, FITTER R. S. R. (2002): Rapid changes in flowering time in British plants.<br />

Science 296:1689–1691.<br />

64


Production of allergenic pollen by ragweed (Ambrosia artemisiifolia L.) is increased in ...<br />

FLEMING DM, SUNDERLAND R, CROSS KW, ROSS AM. 2000. Declining incidence of<br />

episodes of asthma: a study of trends in new episodes presenting to general<br />

practitioners in the period 1989–98. Thorax 55:657–661.<br />

GRIFFITH I. J, POLLOCK J, KLAPPER D. G. (1991): Sequence polymorphism of Amb a I<br />

and Amb a II, the major allergens in Ambrosia artemisiifolia (short ragweed). Int.<br />

Arch. Allergy Appl. Immunol. 96:296-304. Medline.<br />

HOUGHTON J. T, MEIRA L. G, CALLANDER B. A. (1996): Climate Change 1995: The<br />

Science of Climate Change. Cambridge: Cambridge University Press.<br />

HUYNEN M, MENNE B. (2003): Phenology and human health: allergic disorders. Report of<br />

a WHO meeting, Rome, Italy, 16–17 January 2003. Health and global environmental<br />

change, Series No. 1.<br />

JAEGER S, SIWERT N, BERGGREN B. (1996). Trends of some airborne tree pollen in the<br />

Nordic countries and Austria, 1980-1993. Grana: 35:171-178.<br />

KEELING, C. D, WHORF T. P. (2004): Atmospheric carbon dioxide record from Mauna<br />

Loa. Carbon Dioxide Information Analysis Center. Trends: a compendium of data on<br />

global change (online). Last updated 13 February 2004.<br />

http://cdiac.esd.ornl.gov/trends/co2/sio-mlo.htm.<br />

LAIIDI, M., THIBAUDON, M. AND BESANCENOT, J. P (2003): Two statistical<br />

approaches to forecasting the start and duration of the pollen season of Ambrosia in<br />

the area of Lyon (France). Int. J. Biometeorol.<br />

LAST J, GUIDOTTI T. L. (1991): Implications for human health of global ecological<br />

changes. Public Health Rev 1990/91; 18:49–67.<br />

LEWIS W. H, VINAY. P, ZENGER V. E. (1983): Airborne and Allergenic Pollen of North<br />

America. Baltimore, M. D: Johns Hopkins University Press.<br />

MENZEL A. (2000): Trends in phenological phases in Europe between 1951 and 1996. Int.<br />

J. Biometeorol. 44:76–81.<br />

NACLERIO R. M, PROUD D, MOYLAN B. (1997): A double-blind study of the<br />

discontinuation of ragweed immunotherapy. J. Allergy Clin. Immunol. 100:293-300.<br />

Medline.<br />

PAYNE W. W. (1963): The morphology of the inflorescense of ragweeds (Ambrosia-<br />

Franseria: Compositae). Am. J. Bot. 50:872-880.<br />

PETEET D. (2000): Sensitivity and rapidity of vegetational response to abrupt climate<br />

change. Proc Natl. Acad. Sci. USA; 97:1359–61.<br />

PRENTICE I.C, FARQUHAR G D, FASHAM M. J. R. The carbon cycle and atmospheric<br />

carbon dioxide. In: Houghton JT, Ding Y, Griggs DJ, Noguer M, van der Linden PJ,<br />

Dai X, Maskell K, Johnson CA, eds. (2001): Climate change: The scientific basis,<br />

Contribution of Working Group I to the Third Assessment Report of the<br />

Intergovernmental Panel on Climate Change. Cambridge: Cambridge University<br />

Press, 2001; 183–237.<br />

SPIEKSMA F. TH. M, EMBERLIN J. C, HJELMROOS M, JÄGER S, LEUSCHNER R. M.<br />

(1995): Atmospheric birch ( Betula) pollen in Europe: trends and fluctuations in<br />

annual quantities and the starting dates of the seasons. Grana 34:51–57.<br />

WAYNE P, FOSTER S, CONNOLLY J, BAZZAZ F, EPSTEIN P. (2002): Production of<br />

allergenic pollen by ragweed (Ambrosia artemisiifolia L.) is increased in CO 2 -<br />

enriched atmospheres. Ann Allergy Asthma Immunol 2002; 88:279–82.<br />

WEBER R. W. MOTHER (2002): Nature strikes back: global warming, homeostasis, and<br />

implications for allergy. Ann Allergy Asthma Immunol. 88:251–2.<br />

WULFF, R. D, ALEXANDER H. M. (1985): Intraspecific variation in the response to CO 2<br />

enrichment in seeds and seedlings of Plantago lanceolata L. Oecologia; 66:458–60.<br />

WUTHRICH B. (1991): In Switzerland, pollinosis has really increased in the last decade.<br />

Allergy Clin. Immunol. News, 3:41-44.<br />

65


T. Šarić and I. ðalović<br />

ZISKA L, CAULFIELD F. (2000): The potential influence of rising atmospheric carbon<br />

dioxide (CO 2 ) on public health: pollen production of the common ragweed as a test<br />

case. World Res Rev. 12:449-457.<br />

ZISKA L. H, GEBHARD D. E, FRENZ D. A, FAULKNER S, SINGER B. D, STRAKA J.<br />

G. (2003): Cities as harbingers of climate change: common ragweed, urbanization,<br />

and public health. J. Allergy Clin. Immunol. 111:290–5.<br />

66


Herbologia Vol. 7, No. 1, 2006.<br />

GENETICALLY MODIFIED HERBICIDE–TOLERANT CROPS<br />

– STATE AND PERSPECTIVES<br />

Goran Malidža 1 , Vaskrsija Janjić 2 , Ivica ðalović 3<br />

1 Institute of Field and Vegetable Crops – Novi Sad, Serbia and Montenegro<br />

2 Institute of Agricultural Research “Serbia”<br />

Centre for Pesticides and Environment Protection – Zemun, Serbia and Montenegro<br />

3 Faculty of Agronomy – Čačak, Serbia and Montenegro<br />

Abstract<br />

Development and production of genetically modified crops is the<br />

hallmark of the end of last and beginning of new century. The most<br />

remarkable commercial success regarding genetically modified crops has<br />

been achieved with herbicide – tolerant crops as HTCs offer the potential for<br />

many benefits: simpler weed control, more effective management of<br />

problematic and resistant weeds, control of parasitic weeds, use of minimum<br />

additional tool in integrated weed management, avoidance of yieldtillage,<br />

loss caused by current herbicides, etc. Potential risks associated with HTCs<br />

include: gene flow, herbicide resistant volunteers, selection of weed flora in<br />

favour of species less susceptible to herbicides, potential development of<br />

herbicide-resistant weeds, grower's increased dependency on herbicides,<br />

reduced application of integrated weed management, losing of traditional<br />

skills of weed possible decrease in biodiversity in fields, etc.management,<br />

Key words: genetically modified plants, herbicides, tolerance, weeds, resistance, gene flow,<br />

integrated weed control.<br />

Introduction<br />

A significant increase in crop production over the recent decades has<br />

resulted from growing highly improved cultivars, i.e. their grown hybrids as<br />

well as from modern agrotechnical practices.<br />

The latest achievements in molecular genetics, biochemistry and<br />

physiology largely contributed to breeding crops of improved qualites, of<br />

which the tolerance of herbicides seems the most intriguing. Therefore, the<br />

first generation of the genetically modified crops relate to crop inputs or<br />

agronomical properties. Genetically modified herbicide–tolerant crops have<br />

aroused an interest of numerous stakeholders, unanimously sharing an<br />

opinion that farmers, herbicide and plant seed producers benefit from such<br />

crops most, with end–consumers benefiting none (McHugen., 2000; James,<br />

2001).


G. Malidža et al.<br />

Over breeding the herbicide–tolerant crops, the methods of<br />

recombinant DNA (genetic engineering) may be used to obtain genetically<br />

modified, transgenic crops. Biotechnological boom also allowed new<br />

organisms to be bred, their isolation mode to come into being as well as some<br />

of the DNA fragments to be transeferred, and prokaryotic and eukaryotic<br />

genes to be combined, too. Therefore, an ever-lasting dream of geneticists to<br />

combine various pedigree genes when breeding organisms with either<br />

modified or by introducing entirely new traits or entirely new organisms<br />

came into being. (Bekavac et al., 2004).<br />

So far, the combat against weeds has been based on the indigenously<br />

herbicide-tolerant crops. However, so applied herbicides over several decades<br />

so far have led to numerous problems requiring new approaches to be made.<br />

The development of molecular genetics paved the way for genetic<br />

transforming of the individual plant species as well as for expanding genetic<br />

variability and breeding the herbicide– and other toxic compounds–tolerant<br />

crop genotypes. So handled, the genes have encouraged an entirely new<br />

herbicide application technology. The concept of genetic transformation or<br />

genetic engineering embraces all the DNA processes to be transferred from<br />

one organism to another as well as their expression in the host plant. Genetic<br />

transformation is considered to be the procedure allowing genes to be<br />

transformed between the species irrespective of their being cognate and their<br />

number of chromosomes (Hull et al., 2000).<br />

As a technique of foreign DNA transfer to the plant cell, genetic<br />

transformation requires the following:<br />

• identifying the gene to be transferred;<br />

• isolating and cloning the gene;<br />

• introducing the gene into the host genome;<br />

• replicating genetic material;<br />

• expressing genetic material;<br />

• transferring the cell morphogenetic ability, and<br />

• introducing structural genes so as to be inherited through generative<br />

reproduction with safety.<br />

Methodologically, genetic transformation may be attained by directly<br />

introduced structural genes into the host crop genome or through vectors. The<br />

vectors appearing over the gene transfer may be Agrobacterium plasmides,<br />

DNA of the crop viruses, DNA of the plant organisms, DNA of pollen as<br />

well as that of pollen tube. When breeding herbicide-tolerant crops, the<br />

methods of DNA recombinants (genetic engineering) may be attempted, with<br />

virtually genetically modified i.e. transgenic crops. In addition, crop<br />

tolerance towards herbicides may be exerted using somaclonal variabilities,<br />

mutations and commonly used modes in plant breeding (Dyer, 1996., cit. By<br />

Malidza et al., 2005).<br />

68


Genetically modified herbicide-tolerant crops – state and perspectives<br />

Over herbicide–resistant crop growing, the three principles have been<br />

taken into account, as follows:<br />

• introducing the genes responsible for hyper-reproduction of the<br />

enzyme affected by herbicide;<br />

• altering susceptibility of the crucial locus exposed to herbicide impact<br />

and;<br />

• introducing the gene responsible for detoxifying herbicides existing in<br />

the crops.<br />

The largest number of herbicide – resistant crops could be got altering<br />

the main locus exposed to a herbicide, using induced mutations in crops and<br />

microorganisms and introducing microorganism genes to synthesize enzymes<br />

responsible for detoxifying herbicides.<br />

An increasing growth of transgenic herbicide–resistant crops has<br />

largely contributed to weed control in recent years, with an array of merits<br />

favouring producers rather than conventional modes do.<br />

Current issues relating to this field, the significance of the genetically<br />

modified herbicide–resistant crops, their most important traits, a detailed<br />

account on the relevant accomplishments made worldwide so far as well as<br />

possible undesirable effects of the technology considered will be discussed in<br />

the paper.<br />

Genetically modified and herbicide–tolerant crops as a global challenge<br />

Herbicide–tolerant crops do not seem to be a new phenomenon, their<br />

acceptable level resistance being a fundamental precondition of using<br />

herbicides with safety. In addition, a new herbicide is approached by testing<br />

its selectivity compared to that in the leading grown plant species. However,<br />

due to other requirements to be met (toxicological, ecotoxicological, impact–<br />

range, prices etc.) developing new herbicides has been visibly slackened and<br />

become rather costly in recent years. Moreover, grown crops have had their<br />

tolerance altered due to a booming biotechnology, with tremendously<br />

increasing number of herbicide producers making a huge profit. With<br />

herbicide – tolerant crops, lower losses in yield, lower pesticide and<br />

production costs are expected, too. Booming growth of the genetically<br />

modified herbicide – tolerant and insect-resistant crops has upgraded the<br />

conventional crop management since the last decade of the last and the first<br />

one of current century, their canopy highly expanding (Tab.1).<br />

Tab. 1. Global area of transgenic crops from 1996 to 2003 (James, 2003)<br />

Year 1996 1997 1998 1999 2000 2001 2002 2003<br />

Million<br />

1.7 11.0 27.8 39.9 44.2 52.6 58.7 67.7<br />

hectares<br />

69


G. Malidža et al.<br />

As can be seen from Table 1., the total areas under the genetically<br />

modified crops increased by 40 times more from 1996–2003 than those in<br />

previous years. (James, 2003). Thus, in 2003, 73% accounted for total areas<br />

under transgenic crops, of which soybean, canola, maize and cotton did most<br />

(Tab.2).<br />

Table 2. Global area of transgenic herbicide-tolerant crops from 1999 to 2003<br />

(James, 1999, 2000, 2001, 2002, 2003)<br />

Crop<br />

Million hectares<br />

1999 2000 2001 2002 2003<br />

Herbicide tolerant soybean 21,6 25,8 33,3 36,5 41,4<br />

Herbicide–tolerant canola 3,5 2,8 2,7 3 3,6<br />

Herbicide–tolerant maize 1,5 2,1 2,1 2,5 3,2<br />

Bt /Herbicide–tolerant maize 2,1 1,4 1,8 2,2 3,2<br />

Herbicide–tolerant cotton 1,6 2,1 2,5 2,2 1,5<br />

Bt /Herbicide–tolerant cotton 0,8 1,7 2,4 2,2 2,6<br />

*Bt –insect resistance<br />

From the short – term point of view, not believing in no risks of<br />

transgenic crops, the world public seem suspicious and reluctant to put them<br />

into practice. For example, the North and South America countries rank best<br />

by genetically modified produces, the EU countries still lingering about<br />

whether or not to back up such produces. Nonetheless, in Europe, such crops<br />

have been estimated to be notably advantageous. Thus, Phipps and Park<br />

(2002) estimated that in case of growing genetically modified maize,<br />

soybean, canola and cotton resistant to insects, the annual pesticide<br />

consumption in the European Union countries would be reduced by 4.4<br />

million kg, with areas to be treated expected to be reduced by 7.5 million<br />

hectares, which would save roughly 20.5 million litres of petroleum and<br />

reduce carbon – dioxide emission in the atmosphere roughly by 73000<br />

tonnes.<br />

Breeding modes for herbicide–tolerant genotypes<br />

Different biotechnological modes applied to a larger number of crops<br />

gave rise to a notably higher crop resistance to herbicides (Tab.3). As the<br />

most reliable way in breeding herbicide–tolerant genotypes is reckoned<br />

selecting from the existing germplasm. A differing tolerance level to the<br />

herbicides was revealed in wheat (Snape et al., 1991) soybean (Fedtke, 1991)<br />

and some of the oriental grasses (Catanzaro, et al., 1993). As regards the<br />

resistance to herbicides, genetic variability may be attempted, recurrent<br />

70


Genetically modified herbicide-tolerant crops – state and perspectives<br />

selection intensified along with conventional crossing modes to obtain new<br />

cultivars, i.e. hybrids. However, this approach has not been widely accepted<br />

by breeders, primarily due to too poor crop genetic variability to withstand<br />

herbicides.<br />

Tab. 3. Methods of obtaining plants resistant to herbicides<br />

(modified according to Duke, 1996)<br />

Plant species Herbicides Methods<br />

Maize<br />

Glufosinate<br />

Imidazolinone<br />

Ciklosydim<br />

PB<br />

C<br />

C<br />

Wheat Glufosynate PB<br />

Soybean<br />

Glyphosat<br />

AT<br />

Sulphonilurea<br />

Sugar beat<br />

Glufosinate<br />

Sulphonilurea<br />

AT<br />

AT<br />

Barley Glufosinate PB<br />

Tobacco<br />

Glufosinate<br />

Glyfosate<br />

Sulphonilurea<br />

2,4–D<br />

AT<br />

AT<br />

AT<br />

AT<br />

Legend<br />

AT–Transfer og agenes by Agrobacterium tumefaciens<br />

PB–Genetic Gun<br />

C–Tissue culture<br />

S–Plant or seed selection<br />

As a usual method for obtaining herbicide–tolerant crops is assumed<br />

using mutagenous chemicals and X–rays used so far for seed treatment and<br />

mutation inducement, thereby helping separate soybean genotypes resistant<br />

to sulphonilurea herbicides. Tissue culture is highly significant to the<br />

herbicide–resistant genotype breeding. In this sense, growing plant cells in its<br />

tissue culture may exert a range of changes in its genetic composition,<br />

including those in gene distribution, fertility level as well as those in<br />

chromosome arrangement. Such a process, called self-cloning variability has<br />

become an important source of variability, which might have been effectively<br />

used in breeding (Scoweroft and Larkin, 1988). Despite a somewhat success<br />

made in this view at first, a larger number of mutants manifested a range of<br />

undesirable effects due to their low ability in this view. In this sense, an<br />

example of tobacco cell culture selection on glyphosate resistance is often<br />

mentioned. However stable regenerant resistance to glyphosate may be, the<br />

crops regenerized could not be used due to partial sterility and undesirable<br />

agro–economical properties (Dyer et al. 1988). On the other hand, an<br />

71


G. Malidža et al.<br />

example in favour of the mode considered may be exposing maize<br />

embriogene callus culture to Imidazolinon, with some of its resistant portions<br />

being isolated from maize culture (Anderson and Georgheson, 1989).<br />

Additionally, several herbicide–resistant lines possessing one of the two<br />

partially separated dominant alleles responsible for differing resistance level<br />

were regenerized (Newhouse et al. 1991). The mode of recurrent crossings<br />

helped breed the lines with preferable agronomical traits, resulting in 14<br />

hybrids resistant to Imidazolinon (Duke, 1996).<br />

Also, as a mode of breeding herbicide–tolerant crops, hybridization<br />

favours those plant species, which may be crossed with weed species or wild<br />

allies possessing genes resistant to a particular herbicide. Crossing the weed<br />

species Brassica campestris resistant to atrazin with several grown species<br />

from the family Brassica has launched a few commercially significant<br />

genotypes resistant to atrazin, too (Beverski et al., 1980). Despite their yield<br />

being reduced by roughly 20% due to undesirable chloroplast mutation, such<br />

genotypes have been observed to inhabit extremely weeded areas where the<br />

conventionally grown varieties would not be yielding economically enough.<br />

Direct gene transfer claims an absolutely new approach to herbicide –<br />

resistant crop breeding. Numerous genotypes have been obtained from the<br />

directly introduced desirable foreign genes into the host plant genome. DNA<br />

hybridization (Southern blotting) (Sambrook et al., 1987), using DNA<br />

polymerase chain reaction (PCR) (Mullis and Faloona, 1987) and enzymatic<br />

activity analysis are generally accepted methods for checking whether or not<br />

foreign genes have been built in in the host cell gene. Theoretically, the<br />

protoplasts and cells inherent to any plant species may be transformed, but<br />

the problem of how to regenerate transformants from the callus obtained<br />

from plant culture is being inevitably encountered. The modes of direct DNA<br />

transfer may be split up into four groups, as follows:<br />

• genetic transformation through Agrobacterium tumefaciens;<br />

• “Gene gun”;<br />

• stimulating endocytosis chemically;<br />

• electroporation, and<br />

• microinjection.<br />

Genetic transformation through Agrobacterium tumefaciens<br />

The bacteria inherent to the family Agrobacterium are gram negative<br />

soil bacillary microorganisms. Agrobacterium is closely related with<br />

representatives of Rhizobium family responsible for nitrogen fixing in the<br />

root nodular bacteria intrinsic to Agrobacterium tumefaciens, whereas<br />

Agrobacterium rhizogenes may bring about an uncontrollable adventitious<br />

root growth (Janjic, 1996).<br />

72


Genetically modified herbicide-tolerant crops – state and perspectives<br />

DNA transferred from a large bacterium Ti plasmide, subsequently<br />

infecting the plant cell, is presumed a key bacterial activity. The DNA<br />

portion being transferred contains a gene responsible for phytohormone<br />

formation causing tumor tissue. The portion considered is mainly transferred<br />

safely to the host plant genome, so that transgenic plant breeding rests on<br />

such an intrinsically more natural transfer system. The gene responsible for<br />

tumor formation is replaced with a desirable gene and its corresponding<br />

vector formed. Gene transfer through Ti cloned carry–overs Agrobacterium<br />

tumefaciens is mostly used for breeding transgenic plants such as tobacco,<br />

petunia and a range of agronomically significant dicotiledons (Lindsey,<br />

1992). However, the gravest and economically significant drawback of<br />

Agrobacterium system is that it cannot help transform monocolitedon plants<br />

from the family Poaceae. Its numerous species include the most dangerous<br />

and resistant weeds (such as Agropyrum repens (L.), Beauv., Sorghum<br />

halepense (L.) Pers. and the like). The efficacy of gene management through<br />

Agrobacterium is also undeveloped in certain economically significant<br />

dicotiledonous plants such as fruit species, eg. peach and sweet cherry.<br />

Indeed, their tissues do not tolerate an infection caused by Agrobacterium or<br />

are likely to be oversensitive to the bacterial toxins, thereby discouraging<br />

effective genetic transformation.<br />

Gene gun<br />

Gene gun helps introduce sped up tungsten or gold particles 1 – 4 µm<br />

by size covered with DNA. Pollen seeds, cell suspension, callus tissues,<br />

endosperm, leaf, scion vegetation cones and the like may be the object of<br />

transformation. This mode has proven the most efficient for obtaining fertile,<br />

transgenic crops of maize, with a somewhat success made in the experiments<br />

with soybean, beans, rice etc. (Cao et al., 1992). Of some of the genetic gun<br />

designs, the commercially most often used one is Kikker's (1993) gun.<br />

Although this method ensures no stable integration of the DNA introduced<br />

into the plant genome, it has helped transform maize with success.<br />

Chemical stimulation of endocytosis<br />

Protoplast may be transformed by directly stimulated endocytosis.<br />

Under particular conditions (high pH, high content of Ca 2+ ), polyehylene<br />

glycol (PEG) and polyvinyl alcohol are likely to reliably exert genetic<br />

transformation frequency from 1 – 5 x 10 -6 . Even a higher degree of genetic<br />

transformation via endocytosis could be attained in cauliflower amounting to<br />

7.6 x 10 -4 (Tanaka et al., 1984). Methodologically, however complex<br />

endocytosis seems to be for implementing genetic transformation, it has<br />

proven a mere routine with rice, being attempted in a larger number of the<br />

laboratories and varieties (Shimamoto et al., 1989).<br />

73


G. Malidža et al.<br />

Electroporation<br />

Electroporation implies a procedure of exposing cell to electric field<br />

with temporary pores taking place on the plasmalemma through which<br />

macromolecules such as DNA enter the cell.<br />

Microinjection<br />

Microinjection is referred to as a macromolecule (DNA, RNA, protein,<br />

virus or organelle) being directly pushed into the plant cell. Carry–overs of<br />

genetic transformation are directly introduced into the cytoplasm or nucleus<br />

intrinsic to the cells being immobilized on a solid medium using glass pipette<br />

tip 0.5–0.1 µm by size.<br />

This method enables introducing macromolecules and chromosomes<br />

with high precision, with the amount of built – in genetic material controlled<br />

with high accuracy, too. The cell wall is not necessary to remove, being<br />

useful to the species with an undeveloped regenerating system from<br />

protoplast.<br />

The most economically significant genetically modified herbicide–<br />

tolerant crops<br />

Theoretically, crops tolerant to all the herbicides may be bred, but only<br />

the economically important plant species and herbicides with more suitable<br />

properties (glyphosate, glufosynate ammonium, sulfonylurea, imidazolinon<br />

cyclohexandion, bromoxinit and the like) deserve mention. To consider<br />

significance of hazards due to such a technology, every single case, i.e. a<br />

grown crop and herbicide itself, to which the crop may exhibit resistance,<br />

should be analyzed.<br />

Glyphosate–tolerant crops<br />

The first commercialised glyphosate–tolerant crops were obtained<br />

introducing the gene for modified enzyme 5–enolpiruvil–shicimat–3–<br />

phosphate synthetase (EPSPS) imparting the biosynthesis of aromatic amino<br />

acids and being the key locus exposed to this herbicide. The gene for EPSPS,<br />

of lower propensity to glyphosate was isolated from Agrobacterium sp.,<br />

strain CP4, enabling tolerance of the majority of the economically significant<br />

plant species (Wells, 1995). The principle of glyphosate detoxification was<br />

brought to bear introducing it into the plant gene from the bacterium<br />

Achromobacter sp., strain LBAA in order to synthesize an enzyme for<br />

glyphosate oxidoreductase (GOX). This enzyme catalyzes the fissure of C –<br />

N bonds existing in glyphosate up to the metabolites with no herbicide<br />

activity felt (Wells, 1995; Padgette et al., 1996). Taken by areas, the<br />

glyphosate–resistant soybean (Roundup Ready Soybeans) is considered the<br />

74


Genetically modified herbicide-tolerant crops – state and perspectives<br />

world leading transgenic crop, commercially used first in 1996, estimated to<br />

have spread on 41.4 million hectares in 2003, accounting for 61% total areas<br />

under transgenic plants. During the first year of the commercially cultivated<br />

glyphosate–tolerant maize, roughly 380.000 hectares were estimated to be<br />

under glyphosate–tolerant maize in the USA (Talpan, 1998). Apart from<br />

soybean and maize, the glyphosate–tolerant cotton and canola are considered<br />

the most important crops, economically. The results of numerous<br />

experiments suggest applying glyphosate to crops of altered tolerance rather<br />

than using standard weed control chemical practices (Moll, 1997).<br />

Glufosinate–ammonium tolerant crops<br />

Glufosinate–ammonium is amino salt of amino acid of phosphinotricin<br />

obtained from tripeptid bialafos (L–phosphonotricil–L–alanyl–alanyn). Its<br />

mechanism of impact based on inhibiting the enzyme, intrinsic to glutamene<br />

synthetase, responsible for synthesis of glutamine acid (Leason et al., 1982),<br />

resulted in worsened protein synthesis and nitrogen metabolizm with raised<br />

ammonia concentration in the plant cell and, therefore, its raised<br />

phytotoxicity. Certain species of the family Streptomyces produce tripeptid<br />

bialafos as well as an enzyme protecting the host plant from detrimental<br />

effect of its own metabolite.<br />

Further, the so–called BAR gene, isolated from Streptomyces<br />

hygroscopicus, is responsible for the tolerance to bialafos, while the gene<br />

isolated from S. viridichromogenes encodes the synthesis of<br />

phosphonotricin–acetyl–transferasis (PAT gene) for detoxifying<br />

phosphinotricin. Both genes are encoding an enzyme for detoxifying<br />

glufosinate–ammonium through acetylization of amino group, the first<br />

metabolite N–acetyl–glyphosinate being found to have no herbicidal activity.<br />

Glufosinate–ammonium – tolerant crops have been attained using BAR or<br />

PAT genes (De Block et al., 1987, Donn et al., 1990 a, b). This is referred to<br />

as a breeding basis with an array of glufosinate–ammonium–tolerant crops,<br />

canola, maize, soybean and sugar beet being the major ones.<br />

Compared with standard herbicides in sugar beet, soybean and spring<br />

canola, glufosinate–ammonium provided better efficiency (Rasche et al.,<br />

1995), Rasche and Gadsby, 1997). While studying weed control potential<br />

with glufosinate–ammonium –tolerant maize under inland conditions,<br />

glufosinate–ammonium was estimated to have achieved an effect being at<br />

least at the level of standard hebicide combinations if not better when<br />

controlling the dominant weeds (Malidza, 2003).<br />

75


G. Malidža et al.<br />

Imidazolinon– and sulfonilurea–tolerant crops<br />

A larger number of crops is deemed intrinsically resistant to a range of<br />

acetolactate synthesis (ALS) inhibitors because their enzymatic system is<br />

able to metabolize these herbicides prior to seriously inhibiting the targeted<br />

enzyme. Obtaining the imidazolinon- and sulfonilurea–tolerant crops is<br />

considered feasible via standard selection modes, coupled with inducing<br />

mutations and direct gene transfer. Two modes were used to obtain<br />

imidazolinon-tolerant maize. First, pollen mutagenesis helped obtain maize<br />

hybrids (the so–called IT) with ALS enzyme being altered, which, as a<br />

characteristic, was controlled with one dominant gene (Greaves et al., 1993).<br />

Second, contrary to the first mode, the selection in tissue culture with no<br />

mutagenous substances used resulted in the highly imidazolinon–resistant<br />

maize (IR) as well as in the crossed sulfonylurea– and triazolopirimidin–<br />

resistant one (Siehl et al., 1996). IT maize hybrid tolerance in the field<br />

conditions was tested using imazetapir in four times higher amount than<br />

required for weed control, without noticing adverse effects on the yield<br />

(Shaner et al., 1996). IT and IR maize hybrids are commercially used with<br />

tendency to grow only IT maize for allowing new genotypes to be bred with<br />

ease. In addition, the difference in withstanding acetolactate synthesis<br />

between the sensitive and imazetapiron-tolerant IT maize hybrid was sevenfold,<br />

with that in IR maize hybrid estimated to be one thousand times higher.<br />

The crucial enzyme in IT maize hybrid showed no tolerance of<br />

chlorsulphuron and flumetsulam, whereas IR maize hybrid withstood their<br />

200 to 2200 times higher dose rates than the sensitive maize hybrid did (Siehl<br />

et al., 1996). Further, the mutagenesis of microspores allowed the<br />

imidazolinon–tolerant spring canola to be obtained, thereby, evolving a new<br />

epoch in weed control not only for this plant species but also for the weeds<br />

from the family Brassicaceae. Also, wheat seems to be highly economically<br />

significant, tolerance of which to imidazolinon was attained through seed<br />

mutagenesis (Shaner et al., 1996). Of the leading grown crops, imidazolinon–<br />

tolerant sunflower has also paved the way for improving weed control using a<br />

wide–range–impact herbicides after crop and weed emergence. A wild<br />

sunflower originating from the USA helped breed the imidazolinon–tolerant<br />

sunflower, having developed resistance after seven years of using imazetapir<br />

on end (Al–Khatib et al., 1998), possessing a key enzyme resistant several<br />

times higher to imidazolinon. The mode of inheritance was partial dominance<br />

(Miller and Al–Khatib, 2000; Jocic et al., 2001), the total tolerance being<br />

attainable only if both hybrid components were homozygous to this trait.<br />

Apart from being resistant to imazamox, sunflower could tolerate imazetapir<br />

and imazapir, but could not sulfonylurea herbicides (Malidya et al., 2000).<br />

Imazamox in the imidazolinon–resistant sunflower has proved to be efficient<br />

in controlling dominant annual broad–leafed and grassy weeds (Malidza et<br />

76


Genetically modified herbicide-tolerant crops – state and perspectives<br />

al., 2002, 2003). Mutant selection ensured a larger number of the<br />

sulfonylurea–tolerant plants (Sebastian et al., 1989; Dekker and Duke, 1995).<br />

The sulfonylurea–tolerant soybean (STS) only registered on the US<br />

market in 1992 has spread and been applied most widely as yet. The STS<br />

crops can simultaneously metabolize herbicide and have higher tolerance to<br />

acetolactat synthetase. The increased rates of chlorimuron–ethyl and<br />

thiphensuphuron–methyl can be used for weed control in STS soybean<br />

(Young, 1997). The active matter of these herbicides may also be applied to<br />

the conventional weed control system in the USA, but in lower rates due to<br />

lower selectivity. The merit of STS soybean is the higher herbicide dose rates<br />

are applied, the higher efficiency is achieved. Moreover, urea and<br />

imidazolinon are considered to have suitable ecotoxicological properties, the<br />

lower rates of which can be used over a longer period, but invariably<br />

effectively protecting from weeds. Limiting factors of using the<br />

imidazolinon- and sulfonylurea–tolerant crops are assumed resistant weed<br />

growth and rather a constrained crop transfer caused by certain herbicides.<br />

Weed resistance develops swiftly and even several years of applying<br />

herbicides are required. Thus, in Lactuca serriola, resistance was exerted<br />

after five (Mallory–Smith et al., 1990) and in Helianthus annuus after seven<br />

years of the unilaterally used herbicide inhibitor ALS–e (Al–Khatib et al.,<br />

1998).<br />

Cycloxidim– and setoxidim tolerant maize<br />

The herbicides from aryloxiphenoxipropionate and cyclohexandion<br />

groups are being used to control annual and perennial narrow–leaved weeds<br />

in broad–leaved crops. The mode of herbicide impact is inhibiting acetil co–<br />

enzyme being carboxilasis in monocotiledonous plant species. Maize<br />

tolerance of setoxidim, cycloxidim and haloxifop was achieved by selecting<br />

mutants in tissue culture via altering sensitivity of the key locus exposed to a<br />

herbicide (Marshall et al., 1992; Somers, 1994). The mode of inheriting this<br />

trait is partial dominance (Parker et al., 1990). Maize hybrids resistant to<br />

setoxidim are being commercially used in the USA (Poast Protected Maize)<br />

and establishing those resistant to cycloxidim is on its way in Europe. The<br />

herbicides used in the cycloxidim–tolerant maize are paving the way for a<br />

more flexible and efficient controlling of Sorghum halepense from rhizome<br />

and grassy weeds in maize (Malidza, 2001).<br />

Bromoxinyl–tolerant cotton<br />

Cotton resistance to bromoxinyl was attained introducing the gene<br />

intrinsic to bacterium Klebsiella ozaenae, encoding synthesis of the enzyme<br />

bromoxinyl, of a specific nitrilasis, responsible for bromoxinyl detoxification<br />

(Stalker et al., 1988, 1994, cit. Dekker and Duke, 1995). The presence of this<br />

77


G. Malidža et al.<br />

gene ensures a high crop resistance to bromoxinyl, withstanding even ten–<br />

fold higher herbicide rates than practiced ones. Importantly, this approach<br />

contributed significantly to broad – leafed weeds control in cotton due to a<br />

shortage of chemical measures and an onset of phototoxicity caused by the<br />

existing herbicides.<br />

The merits of growing genetically modified herbicide–tolerant crops<br />

The glyphosate– and glufosinate–tolerant crops seem the most<br />

promising thanks to a wide–range impact of such herbicides. Also, the crops<br />

resistant to other herbicides (imidazolinon, sulfonylurea, cyclosidim), the<br />

resistance of which was built in through conventional breeding modes are<br />

promising, too. In general, the genetically modified herbicide – tolerant crops<br />

(GMHT) have the following merits: a facilitated and an economically more<br />

suitable weed control; a more effective weed control due to weed<br />

uncontrollability with herbicides used within the conventionally produced<br />

grown crops; herbicide use with higher flexibility; suitability of using<br />

herbicides after emergence allowing for a critical period and weed<br />

detrimental level; additional potential of controlling weeds resistant to other<br />

herbicides and parasitic weeds; potential for achieving higher yields for<br />

increased grown plant tolerance to herbicides; lower hazards to the<br />

environment through putting the ecotoxicologically more favourable<br />

herbicide-tolerant crops into production, and potential of involving<br />

alternative production systems (no–till and the like).<br />

A facilitated and economically more suitable weed control<br />

One of the most emphasized reasons for increasing (Dewar et al.,<br />

(2000) areas under the individual genetically modified herbicide–tolerant<br />

(GMHT) crops is considered a facilitated weed control along with its lower<br />

cost rather than being with alternative weed control measures. Using one<br />

wide–impact-range herbicide on GMHT crops allows controllability over a<br />

large number of weeds without adding any herbicide to extend the existing<br />

impact-range. In most cases, the application of solely one herbicide in<br />

soybean, maize and sugar beet gives if not better, then at least equal effect to<br />

that with combined herbicides in the conventional production. Thus, in the<br />

conventional production system of sugar beet, several herbicides are<br />

combined reiteratedly, while the GMHT sugar beet receives an equal or even<br />

a better effect of glyphosate or glufosinate-ammonium. In addition to being<br />

facilitated, weed control costs have also been cut down as well as those<br />

relating to glyphosate with soybean, maize and sugar beet, being lower than a<br />

much more expensive alternative system of the chemically controlled weeds.<br />

78


Genetically modified herbicide-tolerant crops – state and perspectives<br />

Thus, as suggested by Dewar et al. (2000), the costs of weed control<br />

with sugar beet were estimated to be from 119 and 110 pounds per hectare in<br />

Great Britain in 1998 and in 1999, those of glyphosate and gluphosinate–<br />

ammonium, used in maximal rates for weed control, from 30 to 60 pounds<br />

per hectare. Even though the GMHT seed expenses are higher, their<br />

introducing will, hopefully, enable economically more acceptable weed<br />

control in the future. Furthermore, areal expansion under the glyphosatetolerant<br />

soybean in the USA gave rise to much reduced costs of the<br />

herbicides to be used in the conventional weed control system (Carpenter et<br />

al., 2002).<br />

More effective weed control on account of its uncontrollability with<br />

the herbicides used within mainstream crop production<br />

Among crucial reasons for introducing GMHT crops lies an answer to<br />

the recent problems encountered in weed control, such as resistance of weeds<br />

to dominant herbicides and weeds, belonging to the same family as the grown<br />

crop does, on which the available herbicides have no or poorer effect than<br />

those efficient with GMHT crops. Therefore, introducing GMHT crops will<br />

be promising for controlling a range of perennial and other weeds similar to a<br />

grown crop, rather than lowly efficient existing selective herbicides used in<br />

the mainstream crop production. Dewar et al. (2000) pointed out the<br />

significance of the glyphosate-resistant sugar beet in which self-emerging<br />

potato will be effectively controlled and the number of nematodes in the<br />

succeeding crops reduced at lower costs. As suggested by May (2003), within<br />

a crop, it is easier to control Cirsium arvense as well as other weeds at lower<br />

costs. Also, Sorghum halepense, Asclepias syriaca and other problematic<br />

weeds with the glyphosate-tolerant soybean have been proven controllable<br />

with notable success in the USA (Culpepper et al.; 2000; Pline et al., 2000).<br />

Numerous examples substantiate an effective controllability of problematic<br />

weeds, such as those from the family Brassisaceae in canola (Merker et al,<br />

2004), followed by Xanthium strumarium in sunflower after emergence<br />

(Malidza et al., 2003), Cynodon dactylon in maize (Malidza and Bekavac,<br />

2001) etc. Evidently, weeds could be highly controllable in GMHT crops,<br />

which is illustrated best by the fields under the crops considered.<br />

Herbicide use with higher flexibility<br />

The higher GMHT crop tolerance of herbicides allows their delayed<br />

application since the crops exhibit tolerance to them in later phases of the<br />

grown crop development (glyphosate, setoxidim, cyclosidim, gluphosinate–<br />

ammonium). For example, glyphosate in glyphosate–resistant maize may be<br />

used last of all the available herbicides (Carpenter et al., 2002). As far as<br />

herbicide use timing and crop and weed growing phase are concerned, the<br />

79


G. Malidža et al.<br />

flexibility of glyphosate and glufosinate–ammonium use in the GMHT sugar<br />

beet seems significantly advantageous rather than being with other herbicides<br />

in this crop (Dewar et al., 2000).<br />

Suitability of using herbicides after emergence allowing for a<br />

critical period and weed detrimental treshold<br />

Speaking about crop growing phase, herbicide application in GMHT<br />

plants is likely to be highly flexible, with herbicides applied after emergence<br />

only if meeting economical requirements. Such a lawfulness applies not only<br />

to the conventionally used herbicides, but also to the individual GMHT crops<br />

for potential herbicide use throughout later GMHT crop and weed growing<br />

phases (Hurle, 1998; Martin et al., 2001).<br />

Additional potential of controlling weeds resistant to other<br />

herbicides<br />

Dominant herbicides used in the leading grown plant species embrace<br />

the representatives of triazin, chloracetamid, carbamide, sulphonilurea,<br />

imidazolinon, aryloxiphenoxipropionate, cyclohexandion groups, with a huge<br />

number of the weed resistant biotypes evidenced so far (Heap, 2004).<br />

Considering that glyphosate and gluphosinate–ammonium were not used on<br />

considerable areas in the crop production in the past, they may additionally<br />

help control weed resistance to the remaining herbicides. The potential of<br />

controlling Helianthus annuus resistance to acelolactat synthetase may serve<br />

as an illustration. That soybean has developed its resistance to the herbicides<br />

dominating this weed is a well-known fact in several countries of the USA,<br />

glyphosate more efficiently controlling soybean than all the alternative<br />

herbicides do in the glyphosate–resistant soybean (Allen et al., 2001).<br />

The potential of achieving higher yields for increasing grown plant<br />

tolerance to herbicides<br />

Under stressful environmental conditions, herbicides cause no adverse<br />

effects in the GMHT crops unlike frequent ones relating to the<br />

conventionally produced crops (Burnside, 1996). With the imidazolinone and<br />

sulphonilurea herbicide groups resistant crops, the risk of extended adverse<br />

effects caused by more persistent herbicide representatives to the succeeding<br />

crops may be reduced over their setting (the imidazolinone-resistant canola,<br />

maize and sunflower as well as the sulphonilurea–resistant soybean and the<br />

like).<br />

The potential of parasitic weeds control<br />

The parasitic weeds from the family Orobanche and Striga occupy over<br />

100 million hectares in the African and Mediterranean countries, thereby<br />

80


Genetically modified herbicide-tolerant crops – state and perspectives<br />

largely constraining the production of a greater range of sensitive grown<br />

crops (Gresset, 2000). Introducing single HT crops and using glyphosate and<br />

acetolactat synthesis inhibitor can visibly reduce the adverse effects of the<br />

weeds considered (Gressel, 1996). More than 100 million farmers in Africa<br />

have been found to loose more than a half of the total maize production due<br />

to weeds from the family Striga (Berner et al., 1995). Growing the<br />

imidazolinone–resistant maize and seed treatment with imazapir confirmed<br />

the potential for controlling parasitic weeds Striga hermontica and S. asiatica<br />

and 3–4 times higher increase in yield (Abayo et al., 1998, Kanampiu et al.,<br />

2003). In addition, introducing the imadadozilinone–resistant sunflower<br />

favoured the control of Orobanche cernua. The potential of concurrent<br />

controlling this parasitic weed and the dominant weeds in the imidazolinoneresistant<br />

sunflower was also substantiated (Malidza et al., 2003).<br />

Lower hazards to the environment by putting the ecotoxicologically<br />

more favourable herbicide-resistant crops into production<br />

Glyphosate, glufosinate–ammonim, imidazolinons, sulphoniluree and<br />

cyclohexandions are considered to have suitable ecotoxicological properties,<br />

which will, when coupled with the previously mentioned ones, contribute to<br />

lower hazards to the environment. As calculated by Wauchope et al. (2001),<br />

replacing atrazin and alachlor with glyphosate and glufosinate–ammonium<br />

when introducing the genetically modified maize toward these herbicides, is<br />

likely to reduce the risk to groundwaters contamination. Since as lower<br />

glyphosate and glufosinate–ammonium dose rates as possible were<br />

attempted, only one fifth up to one tenth of the alachlor and atrazin<br />

concentration was therefore leached into deeper soil layers.<br />

Potential of introducing alternative production system<br />

Tending to use as economically acceptable production system as<br />

possible, the farmers in North America have been massively adopting some<br />

of the GMHT crops, highlighting their suitability for improving no–till<br />

production system. Thus, the areas using no–till production system have been<br />

expanding in the USA since the glufosat–resistant soybean came into life.<br />

The glyphosate – tolerant soybean grown in no–till production mode allowed<br />

weed control in the crop per se, to be more effective, whereas prior to<br />

introducing glyphosate–tolerant soybean, glyphosate had been used before<br />

sowing and emergence for controlling the existing weeds. Within no–till<br />

production system, the glyphosate- and glufosinate–ammonium resistant<br />

maize did not exclude using other herbicides and the best results have been<br />

accomplished combining glyphosate and soil herbicides before sowing with<br />

herbicides used after emergence (Helwig et al., 2003). In addition to no–till<br />

system, the tendency of producing soybean with a higher number of crops per<br />

81


G. Malidža et al.<br />

unit area is increasing. So cultivated glyphosate–resistant soybean requires a<br />

lower number of treatments with herbicides due to different conditions under<br />

which weeds grow (Norsworthy and Oliver, 2001; Reddy, 2003). Also,<br />

introducing some of the GMHT crops will allow herbicides to be used when<br />

cultivating combined crops tolerant of the same herbicide, such as cyclo–<br />

oxide–tolerant beans and maize.<br />

Potenatial hazards to the genetically modified herbicide-tolerant crops<br />

Despite numerous advantages of GMHT crops, encouraging their<br />

cultivation does not necessarily mean a complete absence but inevtiably<br />

present hazards. Increasing dependence of farmers on herbicides is becoming<br />

a serious concern which will either set other measures back or put them out<br />

of practice. Moreover, new mainstream of herbicides and other policies<br />

developed for weed control are expected to be legging behind and ignored.<br />

Importantly, the gravest anxiety related to GMHT crops is associated with<br />

gene transfer to wild allies as well as with resistant weeds growth. Therefore,<br />

due to genetic variability of crops and chemical variability of herbicides, we<br />

cannot generalize hazards. As in previous cases, each case (crop, herbicide,<br />

wild allies and the like) should be regarded per se in light of the potential<br />

hazards and within single cases taking the following into account:<br />

• Which advantage over production of every single crop affords an<br />

additional character of tolerance to a particular herbicide<br />

• What is the likelihood and which are the consequences of gene<br />

transfer responsible for weed and wild allies resistance<br />

• What is the likelihood and which are the consequences of selfemerging<br />

plants or weeds on the rudumentary habitats as the problem<br />

encountered in the agroecosystem<br />

As suggested by previous authors, gene source is unimportant for its<br />

tolerance to herbicide when assessing the hazards considered, i.e. whether<br />

tolerance resulted from mutations within a grown crop or from the gene<br />

coming from another organism. From the other stakeholders’ points of views,<br />

such as consumers and producers, neither gene source nor the herbicide –<br />

tolerant crop mode production is that relevant for such hazards for the time<br />

being. The transfer of genes responsible for tolerance of herbicides may<br />

enhance weeds viability and adaptability to farming and non-farming areas.<br />

In addition, weeds are likely to receive the characters intrinsic to<br />

invading species, which is however more relevant to the genes responsible<br />

for insect resistance or disease causals.<br />

82


Genetically modified herbicide-tolerant crops – state and perspectives<br />

Development of resistant weed biotypes as a result of GMHT<br />

grown areas expansion and of intensified use of lower herbicide number<br />

on larger areas<br />

Weed resistance to herbicides is a genetic phenomenon, being an<br />

example of weed expedited evolution and their intrinsic power to survive.<br />

The process had a long genesis before herbicide–tolerant genetically<br />

modified crops. Expanding GMHT canopies using several wide range–<br />

impact herbicides suggests a faster developing potential of the weed biotypes<br />

resistant to dominant herbicides. 286 resistant biotypes have been recorded in<br />

171 weed species in the world so far (Heap, 2004). An evidence about weed<br />

biotypes resistant to glyphosate in 6 weed species (Heap, 2004) may be<br />

threatening when considering a future crop ranking tolerance to this<br />

herbicide. Lolium rigidum and Eleusine indica developed tolerance to<br />

glyphosate prior to introducing GMHT crops (Powles et al., 1998). Thus, the<br />

first resistant biotype Conyza conadensis was registered in the USA after<br />

three years of glyphosate application in soybean (VanGessel, 2001). More<br />

resistant weed biotypes can be precluded or delayed using combined<br />

herbicides of different impact fashion (Diggle et al., 2003). The same authors<br />

suggest this strategy being more efficient than herbicide rotation of different<br />

exhibiting effect manner. More frequently occurring resistant weed biotypes<br />

over the recent years entail a serious threat to individual GMHT crops,<br />

suggesting their sustainability to be tended over a longer period of time, but<br />

solely as the part of an integrated weed control management (Knezevic and<br />

Cassman, 2003).<br />

Gene transfer from the GMHT to wild allies and weeds<br />

Weed resistance is mainly the onset of selection within a particular<br />

weed population in conditions of reiteratedly used herbicide, where gene<br />

transfer from a grown crop tolerant to a particular herbicide to plant allies is<br />

an additional possibility. Those who deny GMHT crops mainly stress that<br />

such a risk stems from the very centre of origin of grown crops in single<br />

cases. This instance of risk has appeared in maize in central America, but<br />

appears to be insignificant in other areas. Gene transfer is likely to occur only<br />

between the sexually congenial plant species. Keeler et al., (1996) listed only<br />

11 of 60 grown plant species worldwide not to possess any wild allies. Of 13<br />

leading grown plant species, 12 were proven to natuarlly hybridize with wild<br />

cognates (Ellstrand et al., 1999, cit. Wolfenbarger and Phifer, 2000).<br />

Cultivating GMHT crops can increase crossing potential with compatible<br />

species, thereby enhancing the adaptability of the latter to farming and nonfarming<br />

ecosystems. In general, most of the grown crops are regarded unable<br />

to remain viable unless helped by man, with their abilities to survive being<br />

however differing from species to species. Being as much adaptable to the<br />

83


G. Malidža et al.<br />

ecosystems as the herbicide-unresistant crops are, the herbicide–resistant<br />

ones bear no advantages over the other crops in the absence of herbicides<br />

within an ecosystem (Thill, 1996). Further, as a result of gene transfer, a wild<br />

ally of a grown crop is likely to acquire the traits intrinsic to the invading<br />

species. The herbicide–tolerant grown crops do not possess more expressed<br />

traits of the invading species than their non–transgene forms and wild allies<br />

do. Accordingly, direct or indirect impact and control over invading species<br />

have been estimated to roughly 137 billion dollars solely in the USA per<br />

annum (Pimentel et al., 2000, cit. Wolfenbarger and Phifer, 2000). Since<br />

2003, the grown sunflower resistant to imidazolinonim has been cultivated in<br />

the USA. Transfer of the gene, responsible for tolerance of sunflower to<br />

imidazolinonim, to a wild ally was confirmed by Massinga et al., (2003).<br />

Wild sunflower (Helianthus annus) served as a gene donor for<br />

tolerance to imidazolinonim while breeding the grown sunflower, only the<br />

spontaneity of gene transfer in the opposite direction being currently stressed.<br />

Also, that gene for tolerance of wheat (Clearfield) to imidazolinonim may be<br />

transferred to the weed Aegilops cylindrica through hybridization in natural<br />

conditions was revealed by Snyder et al., 2000; Andesron et al., 2004. The<br />

resistance of Aegilops cylindrica in monoculture of the wheat resistant to<br />

imidazolinonim was found by Hanson et al., (2002) to develop in less than 10<br />

years without and in a much shorter time than that with wheat hybridization.<br />

Genetically modified and herbicide-tolerant crops as volunteers in<br />

succeeding crops<br />

This is often emotionally referred to as „super–weed“ by extreme<br />

opponents, emphasizing the risk this technology bears. That grown crops<br />

becoming weeds in the succeeding crops for being linked with each new<br />

herbicide introduced into plant production system, seems nothing of a new<br />

problem. The results obtained in controlling self–fertilized grown plants in<br />

the succeeding ones at the start of growing canola tolerant to glyphosate,<br />

glufosinate–ammonium and imidazolinone appeared to be encouraging.<br />

However, that the spring canola (Brassica napus), being simultaneously<br />

resistant to glyphosate, gluphosinate–ammonium and imidazolinone, had<br />

emerged by spontaneous crossing in Alberta (Canada) was substantiated by<br />

Hall et al. (2000). This is a good example depicting hazards due to an<br />

intermingled resistance toward several herbicides of the congenial foreign–<br />

fertilized crops that might be as problematic as self-fertile crops are in the<br />

ensuing crops. In this case, an integrated crop rotation and herbicide<br />

management along with combining herbicides and other measures for<br />

preventing further progress of such occurrences and adverse effects of self –<br />

fertile crops, B. napus, with multiple resistance to herbicides may be<br />

recommended.<br />

84


Genetically modified herbicide-tolerant crops – state and perspectives<br />

Increasing hazard of damaging untargeted plants using a wide–<br />

impact–range herbicide<br />

Increasingly cultivated GMHT crops have also increased hazards of the<br />

wrongly used herbicides to the crops, not being resistant to a particular<br />

herbicide, and those of the herbicide solution drift on the susceptible ones.<br />

Damages to crops due to wrongly applied herbicides (the application of<br />

herbicide to which crop is not resistant) will occur, but the reasons for this<br />

should not be sought in increasing areas under the genetically modified crops<br />

tolerant to herbicides. Herbicidal drift on untargeted plants appears to be<br />

constant, which, with optimizing water amount for herbicide use, may be<br />

minimized (Ellis et al., 2002).<br />

Potential influence on biodiversity<br />

GMHT crops are gravely threatening biodiversity. If largely grown,<br />

such crops may even change the existing biodiversity of the agroecosystems<br />

subjected to an intense and biased use of the individual cropping practices.<br />

Thus, the relationship between growing genetically modified sugar beet<br />

tolerant to glyphosate and the number of birds, the species Alauda arvensis,<br />

was studied in Great Britain. It is expected that further decline in the number<br />

of this species will depend on a better controllability of Chenopodium album<br />

in the GMHT sugar beet, the seed of which provides this species an important<br />

food source (Watkinson et al., 2000; Dewar et al., 2002, 2003). Of 13 million<br />

hectares, 98% accounted for the glyphosate–tolerant soybean grown in<br />

Argentina in 2003 (James, 2003), its immense growing, giving rise to a fall in<br />

glyphosate prices (3 dollars per litre of glyphosate) expanding its application<br />

to non-farming areas. Over– and recap use of glyphosate (even of up to 16<br />

l/ha/year) caused weeds to entirely disappear not only around the soybean<br />

field and in the soybean crop itself, but also in all the areas treated with<br />

glyphosate being temporarily unused for plant production. In 2000, more than<br />

100 million litres of glyphosate were applied accordingly in Argentina.<br />

Biodiversity is assumed to have largely been jeopardized by an excessively<br />

used glyphosate on the larger areas rather than by the genetically modified<br />

glyphosate–tolerant soybean (Leguaizamon, 2001).<br />

Changes in weed flora<br />

Having created their own systems in weed control, agricultural<br />

producers are using those measures which, in their opinions, render optimal<br />

results. Since weeds adapt to every production system, a good farming<br />

practice may well postpone their adverse effects and the would–be lost<br />

advantages of recent technologies. Unless GMHT crops are used as the part<br />

of integrated weed control system, an expected scenario will be an altered<br />

weed floral composition and growth of its resistant biotypes (Knezevic and<br />

85


G. Malidža et al.<br />

Cassman, 2003). Being so, the best optimality will be ensured by co–<br />

presence of the GMHT crops, the herbicides, to which crops exhibit<br />

tolerance, other herbicides and weed control measures, too. Even though a<br />

wide–impact–range herbicides (glyphosate, gluphosinate-ammonium) are<br />

being applied to the individual, presently, commercialized GMHT crops<br />

worldwide, weed flora is being likely to change by encouraging the naturally<br />

more resistant weeds to such herbicides and by increasing weed sharing, too<br />

(VanGessel, 2001). When introducing individual crops, farmers risk resting,<br />

primarily, on the herbicides, so putting the mainstream weed control<br />

measures out of use.<br />

Prospects and future of the genetically modified crops in Europe<br />

The herbicide–tolerant genetically modified crops entail an additional<br />

option to farmers when controlling weeds, which would be appropriately<br />

used only if incorporated into integrated weed control management (Malidza<br />

et al., 2005).<br />

A rapidly increasing area stretch underneath the herbicide–tolerant<br />

genetically modified crops envisages changes to be on their way in Europe<br />

(Agrow, 2003).<br />

That genetically modified crops are advantageous is doubtless.<br />

However, their use should be approached with utmost care because little is<br />

known about possible implications and adverse effects such crops may imply.<br />

The use of transgenic crops for scientific purposes is allowable in the<br />

country, their further ranking hugely depending on the EU one.<br />

Table 4. Forecast of GMHT crops participation in European Union to 2013<br />

(Agrow, 2003)<br />

Crop<br />

First year of<br />

production<br />

% of area planted to<br />

a particular crop in<br />

2008<br />

% of area planted<br />

to a particular crop<br />

in 2013<br />

Maize 2005–2007 10 35–45<br />

Oilseed rape 2006–2008 0–5 20–30<br />

Soybean 2007–2009 0–10 30–40<br />

Sugar beet 2006–2008 5–10 40–50<br />

Wheat 2008–2011 0 15–25<br />

Rice 2007–2009 0–5 30–40<br />

Cotton 2006–2008 5–10 40–50<br />

86


Genetically modified herbicide-tolerant crops – state and perspectives<br />

The issues considered are being discussed at meetings such as Codex<br />

Alimentarius, OECD Seed Schemes, Convention on biological divergence-<br />

CBD, with regulations varying from country to country, from complete<br />

prohibition of GMHT crops, GMO in Algeria to their full liberalization in the<br />

USA and in Argentina and de facto moratorium in EU countries (Masirevic<br />

and Bugarski, 2004). In light of GMHT crops, the EU countries deserve<br />

special mention. The fact that 22,000 ha of maize, i.e. 2,000 ha BT maize had<br />

been sown in Spain and France in 1998, was pressurized by public opinion,<br />

opposing not only the GMHT crops introduction but also the new<br />

transformant experimentation after the initiative had been made by Denmark,<br />

Greece, France and Luxembourg, so bringing them to a standstill (Phipps and<br />

Park, 2002; Bekavac et al., 2004).<br />

However, the moratorium in the EU countries due to application of<br />

GMHT crops was even more tangled when the USA said it would lodge a<br />

complaint to the World Trade Association, accusing the EU of having based<br />

such a moratorium on the entirely unscientific principles, giving no good<br />

reasons for it any longer. Still, denying GMHT crops by EU seems to have<br />

eased up with more optimism shown recently. In 2002, the EU legalized<br />

banning the moratorium, some of its members still hanging on of whether or<br />

not to accept the GMHT crops and pondering their being tightly linked with<br />

numerous factors.<br />

Despite the existing scenarios for the GMHT crop technology<br />

adaptation approaches, GMHT crops are expected to share 10% total EU<br />

agricultural areas in the next 5 years, during which time, GMHT crops ought<br />

to be validated, enlisted, some of their traits–introduced into the<br />

commercially most important varieties and hybrids (Bekavac et al., 2004).<br />

Overall, in the next ten years, the GMHT crops are expected to share<br />

EU areas, but depending on production specificities from region to region<br />

(weeds and pests) as well as on those from crop to crop (BioPortfolio, 1997–<br />

2003).<br />

References<br />

ABAYO, G.O., ENGLISH, T., EPLEE, R.E., KANAMPIU, F.K., RANSOM, J.K.,<br />

GRESSEL, J. (1998): Control of parasitic witchweeds (Striga spp) on corn (Zea<br />

mays L) resistant to acetolactate synthase inhibitors. Weed Science, 46: 459–466.<br />

AGROW (2003): GM crops growth after EU ban. Agrow, March 28th, PJB Publications,<br />

421: 12.<br />

AL-KHATIB, K., BAUMGARTNER, J.R., PETERSON, D.E., CURRIE, R.S. (1998):<br />

Imazethapyr resistance in common sunflower (Helianthus annuus). Weed Science,<br />

46: 403-407.<br />

ALLEN, J.R., JOHNSON, W.G., SMEDA, R.J., WIEBOLD, W.J., MASSEY, R.E. (2001):<br />

Management of Acetolactate Synthase (ALS)-Resistant Common Sunflower<br />

(Helianthus annuus L.) in Soybean (Glycine max). Weed Technology, 15, 3: 571-<br />

575.<br />

87


G. Malidža et al.<br />

ANDERSON, J.A., MATTHIESEN, L., HEGSTAD, J. (2004): Resistance to an<br />

imidazolinone herbicide is conferred by a gene on chromosome 6DL in the wheat<br />

line cv. 9804. Weed Science, 52, 1: 83–90.<br />

ANNONYMUS (2000): Allergies from GM Food. pp.1–6. http://www. kluyer.stm.tudelft.nl.<br />

BEKAVAC, G., PURAR, BOŽANA, VASIĆ, N., NASTASIĆ, ALEKSANDRA (2004):<br />

Genetički modifikovane biljke–budućnost ili alternativa. Zbornik referata, XXXVIII<br />

Seminar agronoma, pp. 165–172. Naučni institut za ratarstvo i povrtarstvo, Novi Sad.<br />

BERNER, D.K., KLING, B.B., SINGH, B.B. (1995): Striga research and control: a<br />

perspective from Africa. Plant Disease, 79: 652-660.<br />

BEVERSDORF, W. D., WEISS–LERMAN, J., ERICKSON, L. R., AND SOUZA–<br />

MACHADO, V. (1980): Transfer of cytoplasmically–inherited triazine resistence<br />

from birds rape to cultivated Brassica campestris and Brassica napus. Can. J. Genet<br />

Cytol., 22, pp. 167.<br />

BIOPORTFOLIO LTD., 1997 – 2003.<br />

BURNSIDE, O.C. (1996): An agriculturalists viewpoint of risks and benefits of herbicideresistant<br />

cultivars. In: Herbicide Resistant Crops, edited by Duke, S.O., CRC Press,<br />

Inc., 391-406.<br />

CAO, J. DUAN, X., MCELROY, D., WU, R. (1992): Regeneration of herbicide resistant<br />

transgenic rice plants following microprojectile–mediated transformation of<br />

suspension culture cells. Plant Cell Rpts. 11., 586–591.<br />

CARPENTER, J., FELSOT, A., GOODE, T., HAMMIG, M., ONSTAD, D., SANKULA, S.<br />

(2002): Comparative environmental impacts of biotechnology-derived and<br />

traditional soybean, corn and cotton crops. Council for Agricultural Science and<br />

Technology, Ames, Iowa.<br />

CATANZARO, C. J., SKROCH, W. A. AND BURTON, J. D. (1993): Resistance of selected<br />

ornamental grasses to graminicides. Weed Technol., 7, pp. 326.<br />

COMAI, L., FACCIOTTI, D., HIATT, R. W., THOMPSON, G., ROSE, R. E., STALKER,<br />

M. D. (1985): Expresion in plants of a mutant aroA gene from Salmonella<br />

typhimurium confers tolerance to glyphosate. Nature 317, 741–744.<br />

CULPEPPER, A.S., YORK, A.C., BATTS, R.B., JENNINGS, K.M. (2000): Weed<br />

Management in Glophosinatee- and Glyphosate-Resistant Soybean (Glycine max).<br />

Weed Technology, 14, 1: 77-88.<br />

CURRAN, W.S., KNAKE, E.L., LIEBL, R.A. (1991): Corn (Zea mays) injury following use<br />

of clomazone, chlorimuron, imazaquin and imazethapyr. Weed Technology, 5: 539.<br />

DE BLOCK, M., BOTTERMANN, J., VANDEWIELE, M., DOCKX, T., THOEN, C.,<br />

GOSSELE, V., MOVA, N.R., THOMPSON, C., VAN MONTAGU, M.,<br />

LEEMANS, J. (1987): Engineering herbicide resistance in plants by expression of a<br />

detoxifying enzyme. EMBO Journal, 6: 2513-2518.<br />

DEKKER, J., DUKE, S.O. (1995): Herbicide-Resistant Crops. Advances in Agronomy, 54:<br />

69-116.<br />

DEWAR, A., MAY, M., PIDGEON, J. D. (2000): GM sugar beet - the present situation.<br />

British Sugar Beet Review, 68, 2: 22-27.<br />

DEWAR, A.M., HAYLOCK, L.A., MAY, M.J., BEANE, J., PERRY, R.N. (2000):<br />

Glyphosate applied to genetically modified herbicide tolerant sugar beet and<br />

‘volunteer’ potatoes reduces populations of potato cyst nematodes and the number<br />

and size of daughter tubers. Annals of Applied Biology 136: 179-187.<br />

DEWAR, A.M., MAY, M., PIDGEON, J. D. (2002): Management of GM herbicide-tolerant<br />

sugar beet for potential environmental benefit to farmland birds. In IACR Annual<br />

Report 2001-2002. Bolton, S., eds, 44-47.<br />

DEWAR, A.M., MAY, M., WOIWOD, I., HAYLOCK, L., CHAMPION, G., GARNER,<br />

B.H., SANDS, R. J. N., QI, A., PIDGEON, J. (2003). A novel approach to the use<br />

88


Genetically modified herbicide-tolerant crops – state and perspectives<br />

of genetically modified herbicide tolerant crops for environmental benefit.<br />

Proceedings of The Royal Society B, 270: 335-340.<br />

DIGGLE, A.J., NEVE, P.B., SMITH, F. P. (2003): Herbicides used in combination can<br />

reduce the probability of herbicide resistance in finite weed populations. Weed<br />

Research, 43: 371-382.<br />

DONN, G., DIRKS, R., ECKES, P. AND UIJTEWAAL, B. (1990a): Transfer and<br />

exppresion of modified phosphinothricin–acetyltransferase gene from Streptomyces<br />

virido chromogenes in tomato, melon, carrot and strawberry. VII International<br />

Congress of Plant Tissue and Cell Culture. Amsterdam, abstract, 176.<br />

DONN, G., NILGES, M., MOROCZ, S. (1990b): Stable transformation of maize with a<br />

chimaeric modifed phosphinothricin-acetyltransferase gene from Streptomyces<br />

viridochromogenes. In: Abstract VII International Congress of Plant Tissue and Cell<br />

Culture, Amsterdam, 53.<br />

DOTRAY, P. A., MARSHALL, L. C., PARKER, W., P. WYSW, D. L. SOMERS, D. A.<br />

AND GENGENBACH, B. G. (1993): Herbicide tolerance and weed control in<br />

sethoxydim–tolerant corn (Zea mays). Weed Sci. 41. pp. 213.<br />

DUVICK, D. N. (1996): Seed company perspectives: In Herbicide Resistant Crops. (Duke.<br />

S. O. ed.), CRC Press, Inc. New York, pp. 253–262.<br />

DYER, W. E. (1996): Techniques for producing herbicide-resistant crops. In: Herbicide<br />

Resistant Crops, edited by Duke, S.O., CRC Press, Inc., 37-51.<br />

ELLIS, J.M., GRIFFIN, J.L., JONES, C.A. (2002): Effect of Carrier Volume on Corn (Zea<br />

mays) and Soybean (Glycine max) Response to Simulated Drift of Glyphosate and<br />

Glufosinate. Weed Technology, 16, 3: 587-592.<br />

FEDTKE, C. (1991): Deamination of metribuzin in tolerant and susceptible soybean<br />

(Glycine max.) cultivars. Pestic. Sci., 31, pp. 175.<br />

GREAVES, J.A., RUFENER, G.K., CHANG, M.T., KOEHLER, P.H. (1993): Development<br />

of resistance to Pursuite herbicide in corn – The IT gene. Proc. of the 48th Annual<br />

Corn and Sorghum Industry Research Conference, Chicago, 9-10 Dec., 104-118.<br />

GRESSEL, J. (1996): The potential roles fro herbicide-resistant crops in world agriculture.<br />

In: Herbicide Resistant Crops, edited by Duke, S.O., CRC Press, Inc., 231-250.<br />

GRESSEL, J., ROTTEVEEL, T. (2000): Genetic and ecological risks from<br />

biotechnologically derived herbicide-resistant crops: Decision trees for risk<br />

assessment. Plant Breeding Reviews, 18: 251-303.<br />

HALL, L., TOPINKA, K., HUFFMAN, J., DAVIS, L., GOOD, A. (2000): Pollen flow<br />

between herbicide-resistant Brassica napus is the cause of multiple-resistant B.<br />

napus volunteers. Weed Science, 48: 688-694.<br />

HANSON, D.E., BALL, D.A., MALLORY-SMITH, C.A. (2002): Herbicide Resistance in<br />

Jointed Goatgrass (Aegilops cylindrica): Simulated Responses to Agronomic<br />

Practices. Weed Technology, 16, 1: 156-163.<br />

HEAP, I. M. (2004): International survey of herbicide resistant weeds. Online Internet.<br />

Accessed on 10 April 2004. Available: www.weedscience.com.<br />

HELLWIG, K.B., JOHNSON, W.G., MASSEY, R.E. (2003): Weed Management and<br />

Economic Returns in No-Tillage Herbicide-Resistant Corn (Zea mays). Weed<br />

Technology, 17, 2: 239-248.<br />

HULL, R., CONVEY, N. S. DALE, P. (2000): Genetically modified plants and the 35S<br />

promoter: assesing the risks and enhancing the debate. 35S debate, John Innes<br />

Centre, Norwich.<br />

HURLE, K. (1998): Present and future developments in weed control – A view from weed<br />

science. Pflanzenschutz - Nachrichten Bayer 51, 2: 109-128.<br />

JAMES, C. (1999): Preview: Global review of commercialized transgenic crops: 1999.<br />

ISAAA Briefs, Briefs No. 12, ISAAA, Ithaca, NY.ž<br />

89


G. Malidža et al.<br />

JAMES, C. (2000): Preview: Global Status of Commercialized Transgenic Crops: 2000.<br />

ISAAA Briefs No. 21, ISAAA: Ithaca, NY.<br />

JAMES, C. (2001): Preview: Global Status of Commercialized Transgenic Crops: 2001.<br />

ISAAA Briefs No. 24, ISAAA: Ithaca, NY.<br />

JAMES, C. (2002): Preview: Global Status of Commercialized Transgenic Crops: 2002.<br />

ISAAA Briefs No. 27, ISAAA: Ithaca, NY.<br />

JAMES, C. (2003): Preview: Global Status of Commercialized Transgenic Crops: 2003.<br />

ISAAA Briefs No. 30, ISAAA: Ithaca, NY.<br />

JANJIĆ, V. (1996): Savremena istraživanja prirode i delovanja herbicida. Zbornik radova<br />

Petog kongresa o korovima. Banja Koviljača, str. 74–121.<br />

JANJIĆ, V. (2002): Sulphoniluree. Institut za istraživanja u poljoprivredi „Srbija“ i<br />

Akademija nauka i umjetnosti Republike Srpske, beograd, 1–189.<br />

JANJIĆ, V., JOVANOVIĆ, LJ., BLANUŠA, T., MILOŠEVIĆ, D. (2002): Sulphonilurea<br />

herbicides-mode of action. In: Plant Physiology in the New Millenium (eds. S.<br />

Quarrie, B. Krstić and V. Janjić). Institute og Agriculture Research. 101–109.<br />

Belgrade.<br />

JANJIĆ, V:; LJILJANA RADIVOJEVIĆ., SINIŠA MITRIĆ., GORAN MALIDŽA (2004):<br />

Genetičko–biohemijske osnove rezistentnosti korovskih biljaka prema herbicidima<br />

inhibitorima acetolaktat–sintetaze (ALS). Acta herbologica, Vol. 13, No. 2, str.<br />

319–332.<br />

JOCIĆ, S., ŠKORIĆ, D., MALIDŽA, G. (2001): Oplemenjivanje suncokreta na otpornost<br />

prema herbicidima. Zbornik radova Naučnog instituta za ratarstvo i povrtarstvo, 35:<br />

223-233<br />

KANAMPIU, F.K., KABAMBE, V., MASSAWE, C., JASI, L., FRIESEN, D., RANSOM, J.<br />

K., GRESSEL, J. (2003): Multi-site, multi-season field tests demonstrate that<br />

herbicide seed-coating herbicide-resistance maize controls Striga spp. and increases<br />

yields in several African countries. Crop Protection, 22: 697-706.<br />

KEELER, K. H., TURNER, C.E. , BOLICK, M. R. (1996): Movement of crop transgenes<br />

into wild plants. In: Herbicide resistant crops: Agricultural, environmental,<br />

economic, regulatory, and technical aspects, S.O. Duke (ed.), CRC Press, 303-330.<br />

KNEZEVIC, S., CASSMAN, K.G. (2003): Use of herbicide-tolerant crops as a component<br />

of an integrated weed management program. Online, Crop Management,<br />

doi:10.1094/CM-2003-0317-01-MG.<br />

LEASON, M., CUNLIFFE, D., PARKIN, D., LEA, P.J., MIFLIN, B. (1982): Inhibition of<br />

pea leaf glutamine synthetase by metioninsulfoximine, phosphinothricin and other<br />

glutamate analogs. Journal of Phytochemistry, 21: 855-857.<br />

LEGUIZAMON, E. (2001): Transgenic plants in Argentina: present status and implications.<br />

AgBiotechNet, 3: 1-4.<br />

LINDSEY, K. (1992): Genetic manipulation of crop plants. J. Biotechnol., 26, pp. 1–10.<br />

MALIDŽA, G. (2001): Suzbijanje korova u kukuruzu. Zbornik radova Naučnog instituta za<br />

ratarstvo i povrtarstvo, Novi Sad, 71-82.<br />

MALIDŽA, G. (2003): Suzbijanje korova u kukuruzu tolerantnom prema glufosinatamonijumu.<br />

Acta herbologica, 12, 1-2: 67-76.<br />

MALIDŽA, G., BEKAVAC, G. (2001): Suzbijanje korova u transgenom kukuruzu<br />

tolerantnom prema glufosinat-amonijumu i glyphosateu. Prvi meñunarodni<br />

simpozijum Hrana u 21. veku, Subotica, Zbornik rezimea, 193.<br />

MALIDŽA, G., JANJIĆ, V., ðALOVIĆ, I. (2005): Genetički modifikovane biljke tolerantne<br />

prema herbicidima. X Savetovanje o biotehnologiji, Vol. 10. br., 10. str. 96–<br />

112.Agronomski fakultet–Čačak.<br />

MALIDŽA, G., JOCIĆ, S., ŠKORIĆ, D., ORBOVIĆ, B. (2002): Novije mogućnosti<br />

suzbijanja korova u suncokretu. Zbornik radova Naučnog instituta za ratarstvo i<br />

90


Genetically modified herbicide-tolerant crops – state and perspectives<br />

povrtarstvo, 36: 189-205.<br />

MALIDŽA, G., JOCIĆ, ŠKORIĆ, D. (2003): Weed and broomrape (Orobanche cernua)<br />

control in Clearfield sunflower, Proceedings of 7 th Mediterranean Symposium,<br />

Adana, Turkey, 51-52.<br />

MALIDŽA, G., ŠKORIĆ, D., JOCIĆ, S. (2000): Imidazolinone-resistant sunflower<br />

(Helianthus annuus): Inheritance of resistance and response towards selected<br />

sulfonylurea herbicides. Proceedings of 15 th International Sunflower Conference,<br />

Toulouse-France, 42-47.<br />

MALLORY-SMITH, C.A. THILL, D.C., DIAL, M.J. (1990): Development of sulfonylurea<br />

resistant lettuce (Lactuca sativa L.). Weed Sci. Soc. Am. Abstr., 30: 65.<br />

MARSHALL, L-C., SOMERS, D.A., DOTRAY, P.D., GENGENBACH, B.G., WYSE,<br />

D.L., GRONWALD, J. W. (1992): Allelic mutations in acetyl-coenzime A<br />

carboxylase confer herbicide tolerance in maize. Theoretical and Applied Genetics,<br />

83: 435.<br />

MARTIN, S. G., VAN ACKER, R. C., FRIESEN, L. F. (2001): Critical period of weed<br />

control in spring canola. Weed Science, 49, 3: 326-333.<br />

MAŠIREVIĆ, S., BUGARSKI, B. (2004): Genetski modifikovane biljke i semenarstvo.<br />

Zbornik referata, XXXVIII Seminar agronoma, pp. 295–304. Naučni institut za<br />

ratarstvo i povrtarstvo, Novi Sad.<br />

MASSINGA, R. A., AL-KHATIB, K., AMAND, P., MILLER, J. F. (2003): Gene flow from<br />

imidazolinone-resistant domesticated sunflower to wild relatives. Weed Science,<br />

51, 6: 854-862.<br />

MAY, M. J. (2003): Economic consequences for UK farmers of growing GM herbicide<br />

tolerant sugar beet. Annals of Applied Biology, 142: 41-48.<br />

MAZUR, B. J. AND FALCO , C: (1989): The development of herbicide resistant crops.<br />

Ann. Rev. Plant Physiol. Plant Mol. Biol. 40, pp. 441–470.<br />

McHUGEN, A. (2000): Biotechnology and food. American cancel on science and health.<br />

1451–1461.<br />

MERKER, U., RATHKE, G. W., SCHUSTER, C., WARNSTORFF, K., DIEPENBROCK,<br />

W. (2004): Use of glufosinate-ammonium to control cruciferous weed species in<br />

glufosinate-resistant winter oilseed rape, Field Crops Research, 85, 2-3: 237-249.<br />

MILLER, F. J., AL-KHATIB, K. (2000): Development of herbicide resistant germplasm in<br />

sunflower. Proceedings of 15 th International Sunflower Conference, 12-15 June<br />

2000, Toulouse-France, 37-42.<br />

MOLL, S. (1997): Commercial experience and benefits from glyphosate tolerant crops. The<br />

1997 Brighton Crop Protection Conference - Weeds, 931-940.<br />

MULLIS, K. B., AND FALOONA, F. A. (1987): Specific synthesis of DNA in vitro via a<br />

polimerase catalyzed chain reaction. Methods Enzymol., pp. 255–355. In: Herbicide<br />

Resistant Crops., CRC Press, Inc., New York.<br />

NEWHOUSE, K., SINGH, B., SHANER, D. AND STIDHAM, M. (1991): Mutations in<br />

corn (Zea mays L.) conferring resistance to imidazolinone herbicides. Theor. Appl.<br />

Genet. 83, pp. 65.<br />

NORSWORTHY, J. K., OLIVER, L.R. (2001): Effect of Seeding Rate of Drilled<br />

Glyphosate-Resistant Soybean (Glycine max) on Seed Yield and Gross Profit<br />

Margin, Weed Technology, 15, 2: 284-292.<br />

PADGETTE, S.P., RE, D.B., BARRY, G.F., EICHHOLTZ, D.E., DELANNAY, X.,<br />

FUCHS, R.L., KISHORE, G.M., FRALEY, R.T. (1996): New Weed Control<br />

Opportunities: Development of Soybeans with a Roundup Ready Gene. In:<br />

Herbicide Resistant Crops, edited by Duke, S.O., CRC Press, Inc., 53-84.<br />

PARKER, W. B., MARSHALL, L. C., BURTON, J. D., SOMERS, D. A., WYSE, D. L.,<br />

GRONWALD, J. W. (1990): Dominant mutations causing alterations in acetyl-<br />

91


G. Malidža et al.<br />

coenzime A carboxylase confer tolerance to aryloxyphenoxypropionate and<br />

cyclohexanedione herbicides in maize. Proc. Natl. Acad. Sci., U.S.A., 87: 71-75.<br />

PETERSEN, J., HURLE, K. (1998): Einführung von herbizidresistenten Sorten:<br />

Konsequenzen für die Unkrautbekämpfung. Zeitschrift für Pflanzenkrankheiten und<br />

Pflanzenschutz, Sonderheft XVI, 365-372.<br />

PHIPPS, R. H., PARK, J. R. (2002): Environmental benefits of genetically modified crops:<br />

Global and European perspectives on their ability to reduce pesticide use. Journal of<br />

Animal and Feed Sciences, 11: 1-18.<br />

PLINE, W. A., HATZIOS, K. K., HAGOOD, E. S. (2000): Weed and Herbicide-Resistant<br />

Soybean (Glycine max) Response to Glufosinate and Glyphosate Plus Ammonium<br />

Sulfate and Pelargonic Acid. Weed Technology, 14, 4: 667-674.<br />

POWLES, S. B., LORRAINE-COLWILL, D. F., DELLOW, J. J., PRESTON, C. (1998):<br />

Evolved resistance to glyphosate in rigid ryegrass (Lolium rigidum) in Australia.<br />

Weed Science, 46: 604-607.<br />

RASCHE, E., CREMER, J., DONN, G., ZINK, J. (1995): The development of glufosinate<br />

ammonium tolerant crops into the market. The 1995 Brighton Crop Protection<br />

Conference - Weeds, 791-800.<br />

RASCHE, E., GADSBY, M. (1997): Glufosinate ammonium tolerant crops - international<br />

commercial developments and experiences. The 1997 Brighton Crop Protection<br />

Conference - Weeds, 941-946.<br />

REDDY, K. N. (2003): Impact of Rye Cover Crop and Herbicides on Weeds, Yield, and Net<br />

Return in Narrow-Row Transgenic and Conventional Soybean (Glycine max). Weed<br />

Technology, 17, 1: 28-35.<br />

RUDELSHEIM, P. (1993): Engineering crops for tolerance to specific herbicides: a valid<br />

alternative. The brighton Crop Protection Conference–Weeds, Brighton, pp. 265–<br />

272.<br />

SAARI, L. L., COTTERMAN, J. C., PRIMIANI, M. M. (1990): Mechanism of sulfonylurea<br />

herbicide resistance in the broadleaf weed, kochia scoparia. Plant Physiol., 93, 55–<br />

61.<br />

SAMBROOK, J., FRITISCH, E. F. AND MANIATIS, T. (1989): Molecular Cloning–A<br />

laboratory Manual, 2 nd ed., Cold Spring Harbor Laboratory Press. New York, In:<br />

Herbicide Resistant Crops (Duke, S. O., ed.), CRC Press, Inc., New York, pp. 53–<br />

84.<br />

SCOWCROFT, W. R., AND LARKIN, P. J. (1993): Somaclonal variation: Applications of<br />

Plant Cell and Tissue Culture. Wiley, Chichester, U. K. p. 21.<br />

SEBASTIAN, S. A., FADER, G. M., URLICH, J. F., FORNEY, D. R., CHALEFF, R. S.<br />

(1989): Semidominant Soybean Mutation for Resistance to Sulfonylurea<br />

Herbicides. Crop Science, 29: 1403-1408.<br />

SHAH, D. M., HORSCH, R. B., KLEE, H. J., KISHORER, G. M., WINTER, J. A.,<br />

TUMEER, N. E., HIRONAKA, C. M., SANDERS, P. R., GASSER, C. S,<br />

AYKENT, S., SIEGEL, N. R., ROGERS, S. G., FRALEY, R. T. (1986):<br />

Engineering herbicide tolerance in transgenic plants. Science, 223. pp. 478–481.<br />

SHANER, D. L., BASCOMB, N. F., SMITH, W. (1996): Imidazolinone-Resistant Crops:<br />

Selection, Characterization and Managment. In: Herbicide Resistant Crops, edited<br />

by Duke, S.O., CRC Press, Inc., 143-157.<br />

SHIMAMOTO, K., TERADA, R., IZAWA, T., FUJIMOTO, H. (1989): Fertile transgenic<br />

rice plants regenerated from transformed protoplasts. Nature 338, pp. 274–276.<br />

SIEHL, D. L, BENGSTON, A. S., BROCKMAN, J. P., BUTLER, J. H., KRAATZ, G. W.,<br />

LAMOREAUX, R. J., SUBRAMANIAN, M. V. (1996): Patterns of cross-tolerance<br />

to herbicides inhibiting acetohydroxyacid synthase in commercial corn hybrids<br />

designed for tolerance to imidazolinones. Crop Science, 36: 274-278.<br />

92


Genetically modified herbicide-tolerant crops – state and perspectives<br />

SNAPE, J. W., LECKIE, D., PARKER, B. B., AND NEVO, E. (1991): The genetical<br />

analysis and exploatation of differential responses to herbicides in crop secies. In<br />

herbicide Resistance in Weeds and Crops. (Caseley, J. C., Cussans, G. W., and<br />

Atkin, R. K. Eds), Butterworth–Heinemann, Oxford, p. 305.<br />

SNYDER, J. R., MALLORY-SMITH, C. A., BALTER, S., HANSEN, J. L., ZEMETRA, R.<br />

S. (2000): Seed production on Triticum aestivum by Aegilops cylindrica hybrids in<br />

the field. Weed Science, 48, 5: 588-593.<br />

SOMERS, D. A. (1996): Aryloxyphenoxypropionate- and cyclohexanedione-resistant crops.<br />

In: Herbicide Resistant Crops, edited by Duke, S.O., CRC Press, Inc., Boca Raton,<br />

Florida, 175-188.<br />

TALPIN, J. (1998): Transgenic crops: Modest impact upon the markets. Cultivar, november<br />

1998, 15-17.<br />

TANAKA, N., IKEGAMI, M., HOHN, T., MATSUI, C., WATANABE, I. (1984): E. coli<br />

spheroplast–ediated transfer of cauliflower mosaic virus DNA into plant protoplast.<br />

Mol. Gen. Genet. 195, pp. 378–380.<br />

THILL, D. C. (1996): Managing the spread of herbicide resistance. In: Herbicide Resistant<br />

Crops, edited by Duke, S.O., CRC Press, Inc., 331-337.<br />

VANGESSEL, M. J. (2001): Glyphosate-resistant horseweed from Delaware. Weed Science,<br />

49: 703-705.<br />

WANG, Z., ZEMETRA, R. S., HANSEN, J., MALLORY-SMITH, C. A. (2001): The<br />

fertility of wheat × jointed goatgrass hybrid and its backcross progenies. Weed<br />

Science, 49, 3: 340-345.<br />

WATKINSON, A. R., FRECKLETON, R. P., ROBINSON, R. A., SUTHERLAND, W. J.<br />

(2000): Predictions of biodiversity response to genetically modified herbicidetolerant<br />

crops. Science, 289: 1554-1557.<br />

WAUCHOPE, R. D., ESTES, T. L., ALLEN, R., BAKER, J. L., HORNSBY, A.G., JONES,<br />

R.L., RICHARDS, R.P., GUSTAFSON, D. I. (2001): Predicted impact of<br />

transgenic, herbicide-tolerant corn on drinking water quality in vulnerable<br />

watersheds of the mid-western USA. Pest Management Science, 58: 146-160.<br />

WELLS, B. H. (1995): Development of glyphosate tolerant crops into the market. The 1995<br />

Brighton Crop Protection Conference - Weeds, 787-790.<br />

WOLFENBARGER, L.L., PHIFER, P. R. (2000): The ecological risks and benefits of<br />

genetically engineered plants. Science, 290: 2088-2093.<br />

YOUNG, R. D. (1997): The STS seed/herbicide system: A new standard in soybean<br />

production. Proceedings of the Nineteenth Annual Seed Technology Conference,<br />

Iowa State University, Ames, Iowa, 67-70.<br />

93


Herbologia Vol. 7, No. 1, 2006.<br />

Instruction to Authors in Herbologia<br />

One copy of manuscript in English should be submitted by e-mail or<br />

as a hard (paper) copy and a floppy disc.<br />

Manuscripts should be computer typed in MS Word, single spaced,<br />

on the page (paper) format of B5, font of Times New Roman, font size 12<br />

(keywords and list of references with font size 10). The text lines should be<br />

justified. The length of the paper can be up to eight pages.<br />

The paper should start with the title of the article, the names of each<br />

author, his/her institution, address and e-mail address.<br />

Abstract would not exceed 300 words or 20 lines. Keywords, up to<br />

two lines long, should be listed below the abstract.<br />

Main text includes intruduction, materials and methods, results and<br />

discussion. Footnotes should be avoided. SI units should be used. Reference<br />

list should be ordered alphabetically. Examples: AUTHOR, X.Y. & Z.Q.<br />

AUTHOR, 2001: Title of article, Journal title in Italics, 12, 78-84. Or: AUTHOR, A., B.<br />

AUTHOR, 1998: Book title (ed. GH Editor). Publisher, Place, Country.<br />

Figures and tables should be numbered consecutively and should have<br />

an appropriate caption or legend.<br />

Scientific names should be in italic. When a plant name is repeated, it<br />

can be abbreviated, e.g. C. album. For crop plants, common English names<br />

are used, but the scientific name can be given in parentheses at the first<br />

mention in the main text, e.g. oats (Avena sativa). Both British and American<br />

forms of common names can be used (e.g. corn and maize, alfalfa and<br />

lucerne etc.), up to the choice of the author. For herbicides and other<br />

chemicals, in Materials and methods, one should state common approved<br />

names and trade names, e.g. glyphosate (Roundup 360 a.i. L -1 , Monsanto),<br />

and thereafter only trade names. Dose of herbicides should be expressed in<br />

terms of active ingredient (e.g. a.i. ha -1 ).

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!