Articulo 14

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

ANNUAL

REVIEWS Further
Quick links to online content

Ann. Rev. PhylOpalhoi. 1977. 15:165-83


Copyright @ 1977 by Annual Reviews Inc. All rights reserved

APPLICATION OF +3661
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

EPIDEMIOLOGICAL
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

PRINCIPLES TO ACHIEVE
PLANT DISEASE CONTROLI

R. D. Berger
Department of Plant Pathology, University of Florida, Gainesville, Florida 3261 1

INTRODUCTION

Man's early attempts to control the epidemics that plagued his crops were based
largely on methods that evolved through experience, often with little awareness of
the principles governing his success. Our increasing knowledge of the host-patho­
gen-environment disease triangle has enabled us to apply certain principles to reduce
the losses from epidemic disease. The purpose of this chapter is to categorize many
of the epidemiological principles that have recently been put to conscious use in the
control of plant diseases. When possible, particular attention is directed to quantify­
ing the epidemiological response following the application of the technique.
Three epidemiological strategies can be applied to minimize losses due to disease:
(a) eliminate or reduct?the initial inoculum or delay its appearance, (b) slow the
rate of disease increase, and (c) shorten the time of exposure of the crop to the
pathogen. Hundreds of interesting, novel, simple, and complex practices that are
used to control diseases on the world's various crops fit comfortably into one or more
of these categories. The small number of these practices listed here has been chosen
because they pointedly illustrate a particular epidemiological principle.

CONTROL OF DISEASE BY REDUCTION


OF INITIAL INOCULUM

Attempts to reduce initial inoculum have long been of concern to agriculturists in


avoiding losses to epidemics. Most of these procedures come under the general term,

IFlorida Agricultural Experiment Station Journal Series Paper No. 201.


165
166 BERGER

sanitation, which is generally aimed at decreasing the number of propagules of a


pathogen before a crop is planted. A good exposition on various aspects of sanitation
is given by Walker ( 100). Many variations on the general sanitation theme are
available, for example: (a) treatment of seed with hot water or chemicals to kill
seedborne pathogens is used to control black rot and Phoma diseases of cabbage;
(b) seed indexing and certification are used in California and Florida to control
lettuce mosaic virus; (c) citrus budwood registration; (d) potato seed certification;
(e) crop rotation is used to control peanut leaf spot; (j) elimination of alternate hosts
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

frequently is used to control rust diseases; (g) deep plowing of crop refuse is used
to minimize losses to Septoria on wheat; (h) shoot meristem culture (2) is used to
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

eliminate viruses; (i) heat therapy is used to control sugarcane ratoon stunt; and
(j) chemical eradication of pathogen-overseasoning structures is used to control
peach brown rot and apple scab. The vertical resistance concept of Vanderplank (93,
95) has been described as a useful technique to reduce initial inoculum for those
host-pathogen combinations where it applies (21, 34).
Vanderplank (95) analyzed theoretical aspects of epidemics of wheat stem rust
following eradication of barberry. He also calculated the effect on epidemics of
potato late blight following sanitation of cull piles. Often, the barberry and Rhamnus
eradication programs in the United States greatly reduced local outbreaks of wheat
stem rust and oat crown rust, respectively. But the effectiveness of these programs
often was seriously negated by frequent and heavy dispersal of spores up the "Puc­
cinia Pathway" (2 1-23).
Despite the numerous examples where sanitation has been used to reduce initial
inoculum, only a little quantitative information is available on the effectiveness of
sanitation for diseases of the "compound interest" type (95). Sumner & Littrell (91)
showed how epidemics of Helminthosporium maydis on corn were delayed by
several sanitation practices. Kucharek (55) reported several situations where crop
rotation substantially reduced cercospora leafspot on peanuts; however, his conclu­
sions were based on a single disease estimate in mid-season. In another example,
Berger (IS) showed that the progress of an epidemic of Cercospora apii in celery
fields largely depended on the incidence of disease on celery transplants. The effect
of sanitation on soilborne diseases [the "simple interest" diseases of Vanderplank
(95)] has received more attention. Baker and co-workers [references given in (6)]
studied the relation of amount of initial inoculum to subsequent development of
Rhizoctonia damping-off. Wiese & Ravenscroft (104) showed that Cephalosporium
developed less rapidly on wheat when various sanitation practices diminished the
number of surviving fungal propagules.

Sanitation Theory
There is no evidence to show that the rate of epidemic disease progress following
sanitation is the same or different from the rate in comparable crops without
sanitation. Disease progress conceivably could be faster following partial sanitation
if the number of susceptible plants, or the amount of susceptible leaf area, is a
limiting factor in the rate of infection in plots without sanitation. In such cases, the
beneficial effect of the sanitation practice could be rapidly diminished over time.
EPIDEMIOLOGICAL PRINCIPLES FOR CONTROL 167

Obviously, it would be valuable to combine sanitation with practices which decrease


the rate of infection. This combination of procedures has been used successfully to
control early blight of celery (15).
Vanderplank's r (95), the average apparent rate of infection, can be used to
determine the intensity of a given sanitation practice that must be used to limit final
disease below some particular amount. The infection rate (r) is the slope of the linear
regression line determined by plotting loge [x/(l-x)] against time, where (x) is
the proportion of the crop area that is diseased. The equation loge [xo/(l-xo)] =
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

loge [xt/(l-xt)] - (r . t) (Formula I) is used where Xo is the amount of disease


allowed after sanitation to achieve no more than a certain maximum allowable
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

amount of disease (xt) after time (t), assuming an average infection rate (r). For
example, in a crop with a growing season of 80 days and a disease known to have
an r that usually is less than 0. 1 per unit per day, Xo should not exceed 2 X 10-5
if Xt must be kept below 0.05. For (substitution in Formula I):
loge [xo/(l-xo)] = loge (0.05/0.95) - (0. 1 ·80)
= (-2.94) - (8.) = -10.94
and calculation of (xo) gives Xo = 2 X 10-5•
To place this in more realistic terms, if a south Florida celery grower desires no
more than Xt = 0.05 early blight at harvest, he should start the season with fewer
than one Cercospora apii lesion per 1000 celery transplants.
A cautionary statement must quickly be entered here. The calculations above
were based on Vanderplank's notion that epidemics follow a straight logit-line
development (95, 96). Since many epidemics have been shown to have unusually
high (nonlinear) infection rates in initial stages (I, 7, 10, 12, 48-51), the effect of
sanitation may be quickly negated during the early stages of the epidemic. In such
cases it may be necessary to reduce the initial inoculum 100l00-fold more than
required in the calculation of straight logit-line development. That would mean no
more than one lesion per 10,000 to 100,000 transplants using the example calculated
above.
The sanitation principle can be stated in a general manner: The sanitation practice
should reduce the initial inoculum to a sufficiently low level that the normal develop­
ment of disease would not reach a high enough level in the time to harvest to cause
appreciable yield loss (provided unusual influx was avoided).

CONTROL OF DISEASE BY SLOWING THE RATE


OF INFECTION

Most disease control programs are based on techniques to slow the apparent infec­
tion rate of the developing epidemic. Many common control practices, such as
application of pesticides, use of resistant varieties, rouging, etc, fit this category.
Again, innumerable examples are available in the literature, but most fail to quantify
the change in epidemic progress through the utilization of the procedure. I divide
control methods that slow the infection rate into several categories, each of which
is treated separately.
168 BERGER

Effect 0/ Disease Removals on Disease Development


Sanitation practices performed during the development of an epidemic serve not to
reduce initial inoculum, but act as a disease removal mechanism to generally reduce
the substrate on which the pathogen can reproduce. This practice decreases the total
reservoir of secondary inoculum and decreases the rate of infection. Rouging of
infected plants frequently has been used to reduce losses from ins�t-transmitted
diseases.
An outstanding example of the use of sanitation to decrease the rate of infection
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

during an epidemic is the data of Miller, Silverborg & Campana (65) for Dutch elm
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

disease in the city of Syracuse, New York; these data were kindly brought to my
attention by W. Merrill. I have adjusted their data to a 100% base of 53,618 trees
and used a logit transformation to express their results another way (Figure 1). The
sanitation measure, which consisted of removing diseased elm trees, was performed
during the seven-year period 1957-1964. Sanitation allowed the disease to progress
at the rate of r = 0.012 per unit per month. Discontinuance of sanitation measures
in 1964 led to an immediate threefold increase in the rate of disease development
(an average apparent infection rate of r = 0.039 per unit per month). The slow
progress of the disease during the sanitation was likely from escapes, trees with
latent infections that later showed symptoms, and influx of inoculum.

-I

)(
::-3
.......
x

__ OBSERVED

........ PROJECTED

9
-
19h5=3�--��--��----��----��----��--��----��----�'69

Figure J Removal of infected trees (sanitation) slowed the rate of spread of Dutch elm
disease. Average apparent infection rates (r) calculated as per unit per month (95). [Data
adapted from Miller et al (65).] See text for explanation.
EPIDEMIOLOGICAL PRINCIPLES FOR CONTROL 169

If we assume that the original epidemic without sanitation would have progressed
at the same rate as that observed after sanitation ceased (r approximately 0.04
=

per unit per month), then 50% mortality [x = 0.5; loge (x/(1-x)] = 0) would
have been expected by 1963. The seven-year sanitation program delayed the 50%
mortality date five years (1968). If sanitation would have been continued without
interruption, the 50% mortality date would have been reached about 1978 (assum­
ing r 0.012).
=

Van Sickle & Sterner (97) also showed that sanitation resulted in a two- to three­
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

fold decrease in the infection rate of Dutch elm disease in New Brunswick, Canada.
A portion of their data (kindly provided by the authors) was transformed by logits
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

and is presented in Figure 2.


A similar slowing of disease progress with sanitation was observed in the lethal
yellowing disease of coconut palms in Florida (84). Unfortunately, the program to
remove diseased palms was not carried out as stringently as the Dutch elm disease
sanitation program. It was conducted somewhat intermittently, thereby masking
much of the suppressive effect.
Sanitation performed during an epidemic also can be an effective control measure
against leaf spotting organisms. For example, covering a portion of Helmintho­
sporium turcicum infected corn leaves with soil during the final cultivation (lay-by)
-

(.')
o
..J -5

-7

-9�____�______�______�____�______�______�____���
o
YEARS

Figure 2 Sanitation slowed the rate of spread of Dutch elm disease. Infected elm trees
removed only in the Fredericton site. Average apparent infection rates (r) calculated as per
unit per month. [Data provided by Van Sickle & Sterner (97).]
170 BERGER

delayed an epidemic by 5-7 days ( 16). Generally, the earliest infections occurred on
the lower leaves and the lay-by cultivation acted in this case as a disease removal
mechanism.

Reduction in Inoculum Transport and Inoculum Arrival


Restrictions on inoculum transport and arrival by physical or natural means can
also affect epidemic development. If these restrictive measures are applied prior to
the occurrence of disease in the field, they would be considered a sanitation proce­
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

dure. When the same restrictive measures are used to reduce continuing or intermit­
tent influx of inoculum in a developing epidemic, they reduce the infection rate. Few
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

actual examples have appeared in the literature. Nevertheless, this approach re­
mains a fundamental practice in the control of several diseases. For example, the
generally recommended procedure for transplanted crops is to locate seedbeds at
some distance from field plantings to reduce the chance of inoculum influx into the
seedbeds. The use of barrier crops and the interplanting of unrelated host types are
aimed at reducing arriving inoculum. Sequential plantings of a crop in monoculture
increases its vulnerability to disease because inoculum from older plantings fre­
quently threatens the most recent plantings. Some of these control techniques have
found special favor as attempts to reduce movement of vectors for viral diseases (oil
sprays, repellents, aluminum foil, etc). Caution is suggested in the use of wind
barriers, because these sometimes alter the microclimate and may result in increased
disease (33). Fumigation to halt the lateral spread of Dutch elm disease through root
grafts can also be included here. Reducing inoculum transport within a crop fre­
quently is used as a control measure. Avoiding the movement of personnel or'
equipment through fields, particularly when foliage is wet with rain or dew, reduces
the spread of many pathogens. Mechanical topping of plants to restrict the size of
transplants is practiced in the general culture of some crops. If disease is already
present in the plant beds, this topping practice should be avoided because it often
results in extensive disease outbreaks from the massive disperal of inoculum.

Interruption of Pathogen LIfe Cycles


Intentional interruption or alteration of a pathogen life or disease cycle is another
means of disease control which works to decrease the rate of infection. Browning
(21) suggests that the eradication programs for barberry and buckthorn force Puc­
cinia graminis and P. coronata to cycle in the uredial stage. In effect this breaks the
life cycle of this pathogen. Royle (79) reports that such an interruption of the disease
cycle evidently provided control for powdery mildew in hop orchards in the United
Kingdom where mechanical cultivation prevented the maturation of c1eistothecia.
Upon the advent of chemical herbicides, the discontinuance of soil disturbance
around the plants allowed the fungus (Sphaerotheca humuli) to complete its life
cycle and become a serious problem.

Chemical Techniques for Slowing the Progress of an Epidemic


Disease increase is slowed by the application of efficacious chemicals. Critical analy­
sis of disease development using fungicides likely began with the work of Barratt
EPIDEMIOLOGICAL PRINCIPLES FOR CONTROL 171

(3) and later followed by Large (60). Vanderplank (94) has given a more recent
analysis. In time and complexity, chemical techniques range from the early topical
applications of bordeaux mixture in vineyards to the recen. advances of injection
of systemic fungicides and antibiotics.
The art and science of timing fungicide applications to achieve the maximum
control effect has received considerable attention over the years (13, 15-17, 43, 46,
52, 71, 81, 88, 89). The techniques of accurately evaluating fungicide effectiveness
deserve mention (40--42, 95). James et al (40, 42) and Vanderplank (95) provide
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

helpful techniques to determine adequate plot size, to reduce and determine the
amount of interplot interference, and to avoid representational errors. James (41)
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

has also provided guidelines for disease assessment. The following are some general
epidemiological criticisms of most published fungicide tests: the disease estimates
are frequently begun too late in the epidemic (date of onset and features of early
epidemic stages are missed); disease estimates are made at too great an interval (the
erratic advance of disease cannot then be correlated to weather); many estimates are
not accurate enough for epidemiological studies (see 45, 90); the latent period of the
pathogen is often disregarded in interpretation; and host growth is often neglected.

SIMULATION OF FUNGICIDE APPLICATION The effect of a fungicide on disease


progress has been simulated by Zadoks (105). The response of an epidemic of a
leaf-spotting pathogen to application of a fungicide differs from the theoretical curve
proposed by Zadoks because enlargement of lesion size, the appearance of delayed
latent infections, and escapes cause substantial increase in amount of disease after
the effect of the control measure begins (Figure 3). Also, actual disease progress
following the loss of fungicidal effectiveness appears to be much faster than depicted
by Zadoks' simulation and the parallel delay in time proposed by Vanderplank (95)
(curve B, Figure 3). A more realistic curve describing the effect of a fungicide is
shown in curve C, Figure 3. The rapid advance of disease following loss of fungicidal
effectiveness usually exceeds the infection rate in comparable plots that are not
sprayed. Over time, this rapid advance of disease greatly reduces the beneficial effect
of the control treatment. When several epidemics in sprayed and unsprayed fields
were compared, disease severity was less in sprayed fields during the period that
fungicidal protection was provided. However, when spray applications ceased, dis­
ease sometimes reached greater severity in the previously sprayed areas than in the
unsprayed areas because of the rapid disease progress following the loss of fungicidal
effectiveness (10; Berger, unpublished). This is depicted in curve D, Figure 3.
The foregoing observations incidate that once spraying for disease control has
begun, the crop should continue to be sprayed, that is, until such time as any rapid
disease progress would no longer significantly affect yield or quality or until there
is reason to believe the disease may be considerably slowed by environment. The
calculation of the area under the disease progress curve for these conditions may
provide a better correlation of disease effect upon yield than would the final disease
percentage (41, 43, 44, 82).
Fungicide sprays applied early in the epidemic when the number of arriving
spores is relatively small may be more effective than sprays applied later. For many
172 BERGER

en
z
o

o
W
II.
Z
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

a::
w
OJ
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

:;;:
::)
z
(!)
o
..J

TIME

Figure 3 Progress of a leaf-spotting disease following a fungicide application. A Unsprayed.


B Theoretical, adapted from Zadoks (105). C Observed response in actual epidemics. CER­
COS simulations gave similar curves (see Figure 4). D Disease severity of sprayed treatment
sometimes surpasses unsprayed. The delay in time from spray application (arrow) to initial
response is one incubation period.

diseases, as the incidence approaches x = 0.05, all vulnerable host tissue is essen­
tially showered with spores so that the slightest break in fungicidal coverage subjects
the tissue to possible infection. The expansion of the numerous existing lesions and
the accompanying physiological deterioration of affected tissues at x > 0.05 com­
monly mask any positive fungicidal effect. It is at this point in the epidemic that most
growers (and many pathologists!) frequently fail to consider the incubation period
of the fungus. A person must wait at least one incubation period to observe any
slowing of the rate of disease increase brought about by an effective fungicide
application.
The epidemic simulator CERCOS (8) aptly depicted the differential disease control
achieved with a fungicide when the fungicidal effect was initiated in various stages
of an epidemic (Figure 4). With environmental conditions held constant and favor­
able for disease development, a simulated spray applied when the observed disease
was very low (x = 0.0003) slowed the epidemic for about seven days. A similar
spray applied when more disease was observed (x = 0.03) slowed the epidemic by
only about three days. The rapid disease progress following loss of fungicide effec­
tiveness was also clearly simulated.
Some systemic fungicides, e.g. benomyl, in addition to preventing spore germina­
tion, penetration, and infection, provide some theraputic effects on existing latent
infections. In CERCOS-simulated epidemics during constant disease-favorable envi­
ronment, 99.5% cure of latent infections which occurred in the previous 6 days (of
EPIDEMIOLOGICAL PRINCIPLES FOR CONTROL 173

4 -.------,

e-e NONE
c--C SPRAY-DAY 10
2- 1:.-6 SPRAY-DAY 35

0-
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

�-2-
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

.::::.
X

�-4-
o
-'

70
DAYS

Figure 4 CERCOS sim ulation offungicide application. Environment held constant at 25°C
day,17°C night, 12 hr leaf wetness per night and no rain. A single simulated spray was applied
when observed disease was x = 0.0003 (day 10) or x = 0.03 (day 35).

a total latent period of 12 days) slowed disease development only slightly. Under
actual field situations a real benefit for the curative effect of a fungicide occurs when
the timing of a protective spray may have been delayed on a day of high disease
hazard.

Slowing the Progress of an Epidemic by Host Resistance


The use of horizontally resistant varieties (93, 95) to slow epidemic progress has long
been a favorite technique with pathologists and plant breeders. It is perhaps the
simplest, and often the most effective means of control available for many diseases.
Numerous examples can be found in the literature (20, 93, 95), and entire books and
lengthy reviews have been written on host resistance (34, 67, 74, 93). The multiline
approach using several genetically similar host lines (isolines) was covered in a
recent review (21).
The various mechanisms of host resistance affect different stages of the pathogen,
or disease cycle. The end result of using horizontal resistance in an epidemic is
174 BERGER

slower disease development on the resistant varieties. There are several ways to
characterize the specific actions of disease resistance. In epidemiological modeling,
host resistance is commonly handled in the following ways: (0) reduction in the total
number of infections; (b) reduction in the rate of lesion expansion; (c) reduction of
pathogen fructification; (d) lengthening of latent or incubation period; (e) reduction
of spore deposition; if) shortening of infectious period; and (g) increase in number
of propagules necessary to establish infection. Each of these effects is examined in
some detail.
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

REDUCTION OF TOTAL NUMBER OF INFECTIONS Most successful cases of


Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

plant resistance are likely to result from fewer infections on the resistant variety
compared to a more susceptible variety under the same conditions. The resistance
ranges from complete immunity (including nonhost) to barely perceptible tolerance.
This type of resistance is frequently altered by climate, host nutrition, and other
predisposing factors. Vanderplank (95) gives numerous examples; see also (14, 18,
22, 23, 30, 34, 57, 58, 61, 62, 69).

REDUCTION OF LESION EXPANSION Reduction in lesion expansion includes

the appearance of only small lesions on resistant plants as well as slowing of lesion
growth over time. The small lesion reaction types in corn as a result of infection by
He/minthosporium spp. (38) are good examples of this class.

REDUCTION OF SPORULATION Some plant resistance is expressed as a reduc­


tion in sporulation by the pathogen despite a lesion size similar to that observed in
susceptible individuals (61, 62). In those cases where host resistance limits lesion
size, fewer spores are formed per lesion on the resistant types compared to the
susceptible types simply because the total lesion surface available for sporulation is
reduced. Reduction of pathogen reproduction in resistant varieties has also been
observed in a bacterial incited disease (4).

LENGTHENING OF LATENT PERIOD Host resistance that operates by increasing


the latency or incubation time of the pathogen has received only scant attention but
this is one of the most effective types of resistance to decrease the infection rate.
When the latent period of the pathogen is extended, the pathogen can complete
fewer cycles of reproduction in a single crop season. This results in less disease at
harvest. Resistance that extends the latent period has been reported for cereal rust
(27) and bacterial soft rot (5).

REDUCTION OF SPORE DEPOSITION There are few reports in the literature


whereby host surface characteristics affect disease progress. Leaf position, orienta­
tion, or size could significantly affect light penetration into the crop canopy and alter
leaf microclimate. These factors also could have a direct effect on the phenomenon
of spore deposition or catch. Leaf surface characteristics (smooth or setiose, waxy
or rough, etc) also could have an effect on catch. The plant would show outward
resistance (reduced disease incidence) simply because of the reduced catch capabili­
ties.
EPIDEMIOLOGICAL PRINCIPLES FOR CONTROL 175

SHORTENING OF INFECTIOUS PERIOD A shortening of the time that infectious


propagules are produced on diseased host tissue (sporulation) also would slow the
progress of an epidemic because of the reduction in the total number of infectious
units over time. A good example of this type of host resistance is the reduced "zone
of sporing" described by Lapwood (57, 58) for Phytophthora infestans on potato.

HIGH INFECTION THRESHOLD In some resistant varieties of plants a minimum


concentration of pathogen propaguJes appears to be necessary for infection. In such
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

cases, the infection rate would be slower on varieties that become infected only at
high spore loads. This type of resistance apparently is partly responsible for "late
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

rusting" (by Puccinia coronata) in oats (63). The concentration of arriving spores
determines the disease incidence. Plants with this type of resistance normally would
have little or no disease even though neighboring susceptible plants could be consid­
erably blighted. This type of resistance may break down under the pressure of a large
influx of inoculum from nearby susceptible individuals. Whether the resistance
brought about from a high infection threshold differs from the resistance of reduced
number of infections discussed above remains to be determined.

SIMULATION OF RESISTANCE Natural host resistance of the horizontal type


could involve one or more mechanisms of the resistance described above. If several
types of resistance were present in a given variety, the resistance would be additive
since the several resistance mechanisms would individually contribute to decreasing
the infection rate. The epidemic simulator CERCOS (8) was used to illustrate
graphically the effects of several resistance types (Figure 5). When catch, spore
production, lesion size, or percentage of successful infections was reduced by 50%,
the infection rate in initial epidemic stages was reduced by about 25%. When two
types of resistance were used simultaneously in the model, the infection rate was
reduced 50% over the susceptible. The most beneficial individual effect (67% reduc­
tion of rate) came from lengthening the latent period (20 days vs 10 days). It was
interesting that all types of resistance failed to decrease the infection rate as the
epidemic progressed beyond x = 0.01. The failure of resistance to decrease the
infection rate when disease incidence was x > 0.01 has been observed in a natural
epidemic (14).

Inhibiting the Progress of Disease by Management of the Environment

Although man can do little to alter the climate in which he lives, several agricultural
practices have been used to alter the microclimate of a crop and coincidentally
influence the development of epidemic disease.
Hallaire et al (36) treat the influence of irrigation and measurement of the micro­
climate in excellent detail. Rotem & Palti (77) reviewed the effects of irrigation on
the increase of disease. Recently, overhead-sprinkler irrigation has been shown to
increase fire blight of apples (87). Leaf or soil moisture does not always increase the
amount of disease. Lapwood and co-workers (56, 59) have shown that irrigation can
reduce common scab of potatoes.
176 BERGER

4 .-----�
----·
0-0 NONE
a-a (RESF,CATCH,SPORES, or LESIZE) x .5
(RES F a LESIZE)x.s
2- <>-<> LATENT PO x 2

0-
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

x
� -2-
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

"-
x

-6-

-10-1----.----,-----.--
'----,,--
I ---,--
I --.--
I -�
o I� 2'0 30 40 50 60 70
DAYS

Figure 5 CERCOS simulation of resistance mechanisms under the same constant environ­
mental conditions as in Figure 4. NONE = no resistance. RESP = resistance function
determining percentage of successful infections. CATCH = number of spores present on
leaves. SPORES = sporulation parameter. LESIZE = lesion size. LATENT period was 10
days for n o resistance. Numbers i n parentheses are average apparent infection rates for
indicated slope and intervaL See text for explanation.

When overhead-sprinkler systems of irrigation are changed over to trickle or drip


systems, it would be interesting to see what changes this makes in the incidence and
prevalence of pathogens that require a lengthy wet period or are dispersed by
splashing water.
One of the easiest ways to decrease losses during an epidemic is to sow or plant
a crop at a time of year that is less favorable for disease. This technique also could
affect disease development in ways already discussed above. Several examples of this
technique have appeared in the literature (73, 100).
In many crops, mechanization has resulted in the standardization of plant and
row spacing. Plant spacing has sometimes influenced the rate of spread of epidemics.
Diseases generally spread faster among plants at closer spacings (9, 24).
Row orientation (compass direction) is frequently overlooked in management
systems. Haas & Bolwyn (35) showed that rows of bean plants running north-south
EPIDEMIOLOGICAL PRINCIPLES FOR CONTROL 177

in southern Ontario, Canada had more disease (Sclerotinia) than in rows planted
in an east-west (E-W) direction. The prevailing winds were mostly westerly provid­
ing more rapid drying conditions in E-W rows. They also claimed that sunlight
penetrated deeper into the canopy of E-W rows. The orientation of rows also
influences inoculum dispersal patterns, particularly in sequentially planted crops.
Disease is usually maximal when crops are harvested. This disturbance can result
in tremendous influx of inoculum to younger crops downwind if fields are oriented
in this manner. Higher infection rates would result from a repeated influx of inocu­
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

lum. Berger (15) recommended that sequential plantings be made in a direction


opposite to prevailing winds so as to reduce losses caused by an influx of inoculum
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

originating from harvesting operations.


The judicious management of greenhouse temperature and ventilation to control
relative humidity has long been a standard recommendation for control of tomato
leaf mold (Cladosporium) and similar foliage pathogens (100).

The Application of Epidemiological Principles in Disease Forecasting


A thorough understanding of the epidemiological principles as they affect specific
diseases aids the pathologist in developing more accurate disease monitoring and
forecasting systems.
Most early disease forecasting techniques involved the correlation of specific
environmental conditions with disease occurrence (15, 47, 68, 76, 80, 102). Usually,
favorable weather critically influenced one or more stages of the pathogen life cycle.
In actuality, the forecast predicts the time of likely onset or rapid increase of disease.
Sometimes forecasts identify periods when pathogens are relatively inactive. It is not
the purpose here to review all disease forecasting literature, as previous reviews (19,
53, 66, 98, 99) more than suffice. However, an update on some recent forecasting
techniques is warranted.

COMPUTER AIDS The use of computers in disease forecasting will likely increase
because of the computer's record-keeping capabilities, mathematical accuracy, and
the speed of response necessary to make the forecasting method current and reliable.
Voluminous amounts of pertinent weather and biological data can be processed
rapidly to achieve regional or local forecasts (37, 85). Plant growth, disease, and
insect simulation models rely almost entirely on computers in the handling of data.
Computers are also used in information delivery systems to make quicker and more
accurate disease management decisions (31).
The potato late blight forecasting systems of Hyre (39) and Wallin (101) found
only limited acceptance among northern US potato growers until the computer­
ized version (Blitecast) (54) was developed. A similar computerized disease fore­
casting and grower advisory system is available for cercospora leafspot on peanuts
(70).

PATHOGEN MONITORING Monitoring of pathogen activity (generally airborne


fungal spores) can provide additional information and increase the forecasting
accuracy, particularly where release of spores is triggered by specific environmental
178 BERGER

conditions or physical disturbance of the crop (15, 28, 32, 72, 78, 83, 92, 102).
Cournoyer (29) and Roelfs et al (75) followed entire epidemics of Puccinia spp.
utilizing spore monitoring. Propagules of soil pathogens have been monitored to
determine potential disease occurrence (86, 103, 104).

REGRESSION ANALYSIS Multiple regression analysis has been used to define as


many as eleven or more independent biological and climatological variables to
describe the progress of an epidemic (25, 26). The analyses are used to predict future
disease intensity and possible crop loss. James et al (44) also used regression analysis
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

to predict losses of potato tubers to late blight.


Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

INFECTION RATE IN DISEASE PREDICTION When the infection rate of a patho­


gen exhibits little variability each year, then the rate can be used to predict future
epidemic development and spray timing schedules after initial disease occurrence.
Infection rate patterns for Helminthosporium turcicum in com sprayed with fungi­
cide were quite similar over several seasons (Figure 6) (12). A satisfactory disease
forecast and corrective spray schedule was developed using this information (13).
A similar approach was used to determine confidence limits for infection rates of
Uromycesphaseoli in snap beans and allowed the prediction of the amount of disease
at harvest (11). For fusiform rust (Cronartiumfusiforme) in Florida, future disease
in pine can be predicted based largely on prevalence of the alternate host (oak) (R.
A. Schmidt, personal communication).

HELMINTHOSPORIUM TURCICUM

...... 1970

-2
...... 1972
_1973

-4

X-
-6
I


x


OIl
.3 -8

-10
5 15 25 30 5 15 25
APRIL MAY

Figure 6 Epidemic progress of Helminthosporium turcicum on maneb-sprayed sweet corn


(Iobelle) for three seasons at Belle Glade, Florida. The consistent pattern of disease develop­
ment from year to year permits a prediction of disease severity at harvest so that corrective
action can be taken.
EPIDEMIOLOGICAL PRINCIPLES FOR CONTROL 179

HOST PHENOLOGY Infection by a pathogen depends on the availability of sus­


ceptible host tissues. Therefore, control measures can be scheduled to coincide with
specific host phenological events. The seasonal development of apple bud and leaf
tissue has long been used to time fungicidal sprays opportunely to control scab
(Venturia inaequa/is). As another example, Merrill & Kistler (64) reported success­
ful control of Endocronartium harknessii on pine with fungicide sprays timed to
correspond with the emergence of needles from the bud sheath. In some row crops,
good protection against leaf-spotting fungi can be achieved by applying fungicide
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

sprays during the period of rapid increase in leaf area (16).


Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

CONTROL OF DISEASE BY SHORTENING TIME


OF EXPOSURE

Time is the third factor that enters into all equations describing growth (initial
amount and rate being the other two). Commercial growers and plant pathologists
seldom consider altering time of exposure as a possible control for disease but,
nevertheless, this can be a most effective measure.
Any practice that shortens the time of exposure of a crop to a pathogen will
decrease the risk of loss due to an epidemic. In a transplanted crop, the setting of
large, vigorous plants with a well-developed root system would insure quick estab­
lishment of the crop and shorten the time to maturity. Use of short season varieties
and maintenance of adequate soil fertility and moisture to avoid any slowdown of
crop growth are the most common techniques to shorten the exposure of crops to
pathogens. In the event an epidemic occurred, the pathogen would have less time
before the disease could reach an economic loss level. To paraphrase one of "Mur­
phy's laws," "the longer a crop remains in the field, the more likely something bad
will happen to it."

CONCLUDING REMARKS

The condensation of the voluminous literature that was available to treat this
chapter's theme was not an easy task. I was especially pleased to have had the
cooperative response from the numerous pathologists who answered my requests for
information. I am certain many individuals can think of additional important con­
trol techniques that I omitted or overlooked. Some may question my choice of
examples for the principles I categorized, particularly if their favorite disease was
neglected. The actual quantification of the epidemiological response to new, or even
long-standing, control practices is frequently neglected. Much work lies ahead. It
was my intention to treat several principles controversially so as to stimulate more
critical analysis in future experimentation. It would not be at all surprising if several
well-established theories on spread of epidemics were eventually set aside as our
knowledge of modern epidemiology increases. New modeling techniques and com­
puter simulation should simplify the assessment of specific epidemiologic events.
These techniques also can aid in determining the underlying biological and ecologi­
cal basis for naturally occurring disease phenomena.
180 BERGER

Literature Cited

1. Analytis, S. 1973. Zur Methodik der 16. Berger, R. D. 1973. Helminthosporium


Analyse von Epidemien dargestellt am turcicum lesion numbers related to
Apfelschorf ( Venturia inaequalis numbers of trapped spores and fungi­
(Cooke) Aderh.) Acta Phytomed. I: cide sprays. Phytopathology 63:930-33
1-76 17. Berger, R. D. 1973. Disease progress of
2. Baker, R., Phillips, D. J. 1962. Obtain­ Cercospora apii resistant to benomyl.
ing pathogen-free stock by shoot tip cul­ Plant Dis. Reptr. 57:837-40
ture. Phytopathology 52:1242-44 18. Berger, R. D., Wolf, E. A. 1968. For­
3. Barratt, R. W. 1945. Intraseasonal ad­
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

eign introductions of Apium as a possi­


vance of disease to evaluate fungicides ble source of resistance to celery bac­
or genetical differences. Phytopathology terial blight (Pseudomonas cichorii).
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

35:654 Phytopathology 58:1043


4. Bartz, J. A., Crill, J. P., John, C. A. 19. Bourke, P. M. A. 1970. Use of weather
1975. Inheritance of tolerance to Er­ information in the prediction of plant
winia carotovora in Florida MH-I disease epiphytotics. Ann. Rev. Phytopa­
tomato. Phytopathology 65:1146-50 thol. 8:345-70
5. Bartz, J. A., Stall, W. M. 1974. Toler­ 20. Browning, J. A. 1974. Relevance of
ance of fruit from different pepper lines knowledge about natural ecosystems to
to Erwinia carotovora. Phytopathology development of pest management pro­
64:1290-93 grams for agro-ecosystems. Proc. Am.
6. Benson, D. M., Baker, R. 1974. Phytopathol. Soc. 1:191-99
Epidemiology of Rhizoctonia solani 21. Browning, J. A., Frey, K. J. 1969. Mul­
preemergence damping-off of radish: tiline cultivars as a means of disease
survival. Phytopathology 64: 1163-68 control. Ann. Rev. PhytopathoL 7:
7. Berger, R. D., Mishoe, J. W. 1976. 355-82
CSMP simulation of several growth 22. Browning, J. A., Simons, M. D., Frey,
functions to describe epidemic progress. K. J., Murphy, H. C. 1969. Regional
Proc. Am. Phytopathol. Soc. 3:217 deployment for conservation of oat
8. Berger, R. D. 1976. Computer simula­ crown-rust resistance genes. Iowa Agric.
tion of Cercospora apii and Helmintho­ Home Econ. Exp. Sm. Spec. Rept.
.
sporium turcicum. Proc. Am. Phytopa­ 64:49-56
thoL Soc. 3:217 23. Browning, J. A., Frey, K. J., Grinde­
9. Berger, R. D. 1975. Disease incidence land, R. L. 1964. Breeding multiline oat
and infection rates of Cercospora apii in varieties for Iowa. Iowa Farm Sci.
plant spacing plots. Phytopathology 18(8) :5-8
65:485-87 24. Burdon, J. J., Chilvers, G. A. 1975.
10. Berger, R. D. 1975. Rapid disease Epidemiology of damping-Off diseases
progress in early epidemic stages. Proc. (Pythium irregulare) in relation to den­
Am. Phytopathol. Soc. 2:35 sity of Lepidium sativum seedlings.
11. Berger, R. D. 1975. Predicting the Ann. Appl. Bioi. 81:135-43
progress of bean rust (Uromyces 25. Burleigh, 1. R., Eversmeyer, M. G., Ro­
phaseo/i) by average apparent infection elfs, A. P. 1972. Development of linear
rates. Pmc. Am. PhytopathoL Soc. 2:73 equations for predicting wheat leaf rust.
12. Berger, R. D. 1974. Prediction of Phytopathology 62:947-53
amount of Helminthosparium turcicum 26. Butt, D. J., Royle, D. J. 1974. Multiple
at harvest in Florida sweet corn. regression analysis in the epidemiology
Phytopathology 64:767 of plant diseases. Epidemics of Plant
13. Berger, R. D. 1974. A guide to spraying Diseases: Mathematical Analysis and
sweet corn for Helminthosporium Modeling. ed. J. Kranz, pp. 78-114.
blight by likely disease spread. Univ. FL New York: Springer. 170 pp.
A gric. Res. Educ. Cent., Belle Glade 27. Clifford, B. C. 1972. The histology of
Res. Rept. EV-1974-15. 3 pp. race non-specific resistance to Puccinia
14. Berger, R. D. 1973. Infection rates of hordei Otth. in barley. Proc. Eur. Medi­
Cercospora apii in mixed populations terr. Cereal Rusts Coni, Prague I:
of susceptible and tolerant celery. 75-78
Phytopathology 63:535-37 28. Cohen, Y., Rotem, J. 1971. Dispersal
15. Berger, R. D. 1973. Early blight of and viability of sporangia of Pse udo ­
celery: analysis of disease spread in peronospora cubensis. Trans. Br. MycoL
Florida. Phytopathology 63:1161-65 Soc. 57:67-74
EPIDEMIOLOGICAL PRINCIPLES FOR CONTROL 18 1

29. Cournoyer, B. M. 1970. Crown rust epi­ potato and loss in tuber yield. Phytopa­
phytology with emphasis on the quantity thology 62:92-96
and periodicity 0/ spore dispersal/rom 44. James, W. c., Shih, C. S., Callbeck, L.
heterogeneous oat cultivar-rust race pop­ C., Hodgson, W. A. 1971. A method for
ulations. PhD thesis. Iowa State Univ., estimating the loss in tuber yield caused
Ames. 191 pp. by late blight of potato. Am. Potato J.
30. Crill, P., Jones, J. P., Burgis, D. S. 1973. 48:457-63
Failure of "horizontal resistance" to 45. Jarvis, W. R., Hawthorne, B. T. 1972.
control fusarium wilt of tomato. Plant Sclerotinia minor on lettuce: progress of
Dis. Reptr. 57:119-21 an epidemic. Ann. Appl. Bioi. 70:207-14
31. Croft, B. A., Howes, J. L., Welch, S. M. 46. Jenkins, J. E. E., Storey, J. F. 1975. In­
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

1976. A computer-based, extension pest fluence of spray timing for the control of
management delivery system. Environ. powdery mildew on the yield of spring
Entomol. 5:20-34 barley. Plant Pathol. 24:125-34
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

32. Eversmeyer, M. G., Kramer, C. L. 47. Jensen, R. E., Boyle, L. W. 1966. A


1975. Air spora above a Kansas wheat technique for forecasting leafspot on
field. Phytopathology 65:490-92 peanuts. Plant Dis. Reptr. 50:810-14
33. Eversmeyer, M. G., Skidmore, E. L. 48. Jowett, D., Browning, J. A., Haning, B.
1974. Wheat leaf and stem rust develop­ C. 1974. Nonlinear disease progress
ment near a wind barrier. Plant Dis. curves. See Ref. 26, pp. 115-36
Reptr. 58:459-63 49. Kranz, J. 1974.. Comparison of epidem­
34. Frey, K. J., Browning, J. A., Simons, ics. Ann. Rev. Phytopathol. 12:355-74
M. D. 1973. Management of host resis­ 50. Kranz, J. 1974. The role and scope of
tance genes to control diseases. Z. mathematical analysis and modeling in
Pjlanzenkr. Pjlanzenschutz 80:160-80 epidemiology. See Ref. 26, pp. 7-54
35. Haas, J. H., Bolwyn, B. 1972. Ecology 51. Kranz, J., Mogk, M., Stumpf, A. 1973.
and epidemiology of Sclerotinia wilt of EPIVEN-ein Simulator fUr Apfel­
white beans in Ontario. Can. J. Plant schorf. Z. Pjlanzenkr. Pjlanzenschutz
Sci. 52:525-33 80:181-87
36. HaJiair, M., Rapilly, F., Pauvert, P. 52. Kranz, J. 1965. Feldversuche zur
1969. Effects de I'irrigation, sous ses Bekiimpfung der Sigatoka-Krankheit
differents modes, sur la biologie l'etiolo­ der Banane (Mycosphaerella musicola
gie et I'epidemiologie des maladies des Leach) in Guinea. Phytopathol. Z.
plantes. Ann. Phytopathol. 1:9-22 52:335-48
37. Harrington, R. B., Giese, R. L. 1974. 53. Krause, R. A., Massie, L. B. 1975. Pre­
MIRACLE- a hierarchial data acquisi­ dictive systems: Modern approaches to
tion system for laboratory automation. disease control. Ann. Rev. Phytopathol.
Comput. Con] Fed. Proc. 33:2393-95 13:31-47
38. Hooker, A. L. 1974. Seedling reactions 54. Krause, R. A., Massie, L. B., Hyre, R.
of corn hybrids to Helminthosporium A. 1975. Blitecast: a computerized fore­
leafspot. Plant Dis. Reptr. 58:975-77 cast of potato late blight. Plant Dis.
39. Hyre, R. A. 1954. Progress in forecast­ Reptr. 59:95-98
ing late blight of potato and tomato. 55. Kucharek, T. A. 1975. Reduction of
Plant Dis. Reptr. 38:245-53 cercospora leafspots of peanut with
40. James, W. C., Shih, C. S., Hodgson, W. crop rotation. Plant Dis. Reptr. 59:
A., Callbeck, L. C. 1976. Representa­ 822-23
tional errors due to interplot interfer­ 56. Lapwood, D. H., Adams, M. J. 1973.
ence in field experiments with late blight The effect of a few days of rain on the
of potato. Phytopathology 66:695-700 distribution of common scab (Strep­
41. James, W. C. 1974. Assessment of plant tomyces scabies) on young potato tu­
diseases and losses. Ann. Rev. Phytopa­ bers. Ann. Appl. Bioi. 73:277-83
thaI. 12:27-48 57. Lapwood, D. H. 1971. Observations on
42. James, W. C., Shih, C. S., CaJibeck, L. blight (Phytophthora in/estans) and re­
C., Hodgson, W. A. 1973. Interplot in­ sistant potatoes at Toluca, Mexico.
terference in field experiments with late Ann. Appl. BioI. 68:41-53
blight of potato (Phytophthora in/es­ 58. Lapwood, D. H. 1963. Potato haulm
tans). Phytopathology 63:1269-75 resistance to Phytophthora in/estans.
43. James, W. C., Shih, C. S., Hodgson, W. IV. Laboratory and field estimates com­
A., CaJibeck, L. C. 1972. The quantita­ pared, and further field analysis. Ann.
tive relationship between late blight of Appl. BioI. 51:17-28
182 BERGER

59. Lapwood, D. H., Wellings, L. W., Haw­ Puccinia striiformis Westend. Ann.
kins, J. H. 1 973. Irrigation as a practical Phytopathol. 2:5-3)
means to control potato common scab 73. Rapilly, F. 1968. Etudes sur l'ergot du
f
(Streptomyces scabies): Final eX eri­ bie: Claviceps purpurea (Fr.) Tul. Ann.
ment and conclusions. Plant Patho . 22: Epiphyties 19:305-29
35-41 74. Roane, C. W. 1973. Trends in breeding
60. Large, E. C. 1952. The interpretation of for disease resistance in crops. Ann.
progress curves for potato blight and Rev. Phytopathol. 1 1 :463-86
other plant diseases. Plant Pathol. 75. Roelfs, A. P., McVey, D. V., Long, D.
1 : 109-17 L., Rowell, J. B. 1 972. Natural rust epi­
61. Leonard, K. J. 1969. Factors affecting demics in wheat nurseries as affected by
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

rates of stem rust increase in mixed inoculum density. Plant Dis. Reptr.
plantings of susceptible and resistant 56:410-14
oat varieties. Phytopathology 59:
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

76. Rotem, J., Ben-Joseph, J. 1 970. Evapo­


1 845-50 ration rate as an indicator for potato
62. Leonard, K. J. 1969. Selection in late blight development in plots of
heterogeneous populations of Puccinia different foliage density. Plant Dis.
graminis f. sp. avenae. Phytopathology Reptr. 54:768-7 1
59: 1 85 1-57 77. Rotem, J., Paiti, J. 1969. Irrigation and
63. Luke, H. H., Chapman, W. H. Barnett, plant diseases. Ann. Rev. Phytopathol.
R. D. 1972. Horizontal resistance of 7:267-88
Red Rustproof oats to crown rust. 78. Rowe, R. c., Johnston, S. A., Beute, M.
Phytopathology 62:414-17 K. 1974. Formation and dispersal of
64. Merrill, W., Kistler, B. R. 1 974. Cylindrocladium crotalariae microscle­
Phenology and control of Endocronar­ rotia in infected peanut roots. Phytopa­
tium harknessii in Pennsylvania. thology 64: 1294-97
Phytopathology 66: 1 246-48
79. Royle, D. J., Griffin, M. J. 1 973. Side­
65. Miller, C., Silverborg, S. B., Cam­
H.
effects of downy mildew fungicides on
pana, R. J. 1969. Dutch elm disease: the incidence of hop powdery mildew
Relation of spread and intensification to
(Sphaerotheca humuli). Plant Pathol.
control by sanitation in Syracuse, New
22: 1 29-33
York. Plant Dis. Reptr. 53:551-55
80. Royle, D. J. 1970. Infection periods in
66. Miller, P. R, 1959. Plant disease fore­ relation to the natural development of
casting. Plant Pathology Problems and
hop downy mildew (Pseudoperonospora
Progress, pp. 557-65. Madison: Univ.
humuli) Ann. Appl. BioI. 66:281-91
Wis. Press. 588 pp.
8 1 . Schenck, N. C. 1968. Fungicidal con­
67. Nelson, R. R., ed. 1 973. Breeding plants
for disease resistance, concepts and ap­
trol of watermelon downy mildew and
its relationship to first infection in the
plications. University Park: Penn. State
Univ. Press. 401 pp. field. Plant Dis. Reptr. 52:979-80
68. Palti, J., Rotem, J. 1 973. Epidemiologi­ 82. Schneider, R. W., Williams, R. J., Sin­
cal limitations to the forecasting of clair, J. B. 1976. Cercospora leafspot of
downy mildews and late blight in Israel. cowpea: models for estimating yield
Phytoparasitica 1 : 1 1 9-26 loss. Phytopathology 66:384-88
69. Parlevliet, J. E. 1976. Evaluation of the 83. Schrodter, H. 1960. Dispersal by air
concept of horizontal resistance in the and water-the flight and landing.
barley/Puccinia hordei host-pathogen Plant Pathology, An Advanced Treatise,
relationship. Phytopathology 66:494-97 ed. J. G. Horsfall, A. E. Dimond,
70. Parvin, D. W., Smith, D. H., Crosby, F. 3 : 1 69-227. New York: Academic
L. 1 974. Development and evaluation of 84. Seymour, C. P. 1974. Current status of
a computerized forecasting method the coconut lethal yellowing problem in
for Cercospora leafspot of peanuts. Florida. Proc. Tall Timbers Con! Ecol.
Phytopathology 64:385-88 Anim. Control Habitat Manage. 6:
7 1 . Rapilly, F. 1970. La determination des 1 2 1-23
dates de traitments fongicides appliques 85. Shaner, G. E., Peart, R. M., Newman,
par voie aerienne sur cereales in vegeta­ J. E., Stirm, W. L., Loewer, O. L. 1 972.
tion, cas du ble d'hiver. Phytiatr.­ EPIMAY-an evaluation of a plant dis­
Phytopharm. 19: 1 85-203 ease display model. Purdue Univ. Res.
72. Rapilly, F; , Fournet, J., Skajennikoff, Bull. 890. 15 pp.
M. 1 970. Etudes sur l'epidemiologie et 86. Sherwood, R. T., Hagedorn, D. J. 1958.
la biologie de la rouille jaune du ble Determining the common root rot po-
EPIDEMIOLOGICAL PRINCIPLES FOR CONTROL 1 83

tential of pea fields. Wis. Agric. Exp. 96. Vanderplank, J. E. 1 960. Analysis of
Sm. BulL 531. 12 pp. epidemics. See Ref. 83, pp. 229-89
87. Spotts, R. A., Stang, E. J., Ferree, D. C. 97. Van Sickle, G. A., Sterner, T. E. 1 976.
1976. Effect of overtree misting for Sanitation: a practical protection
bloom delay on incidence of fire blight against Dutch elm disease in Frederic­
in apple. Plant Dis. Reptr. 60:329-30 ton, New Brunswick. Plant Dis. R eptr.
88. Steiner, K. G. 1973. The influence of 60:336-38
fungicidal treatment on the develop­ 98. Waggoner, P. E. 1 965. Microclimate
ment of coffee berry disease (Colletotri­ and plant disease. Ann. Rev. Phytopa­
chum coffeanum NOACK). Z Pjlan­ thol. 3:103-26
zenkr. Pjlanzenschutz 80:671 -8 1 99. Waggoner, P. E. 1 960. Forecasting epi­
by Moscow State University - Scientific Library of Lomonosov on 11/04/13. For personal use only.

89. Steiner, K. G . 1 973. Duration o f the demics. See Ref. 83, pp. 291-3 1 2
activity of fungicides against coffee 100. Walker, J . C . 1957. Plant Pathology.
berry disease ( Colletotrichum cojfea­ New York: McGraw-Hill. 707 pp.
Annu. Rev. Phytopathol. 1977.15:165-181. Downloaded from www.annualreviews.org

num NOACK). Z Pjlanzenkr. Pjlan­ 2nd ed.


zenschutz 80:532-46 1 0 1 . Wallin, J. R. 1 962. Summary of recent
90. Strandberg, J. 1973. Spatial distribution progress in predicting late blight epi­
of cabbage black rot and the estimation demics in United States and Canada.
of diseased plant populations. Phytopa­ Am. Potato J. 39:306-12
thology 63:998-1003 102. Watson, M. A., Heathcote, G. D.,
9 1 . Sumner, D. R., Littrell, R. H. 1 974. In­ Lauckner, F. B., Sowray, P. A. 1975.
fluence of tillage, planting date, inocu­ The use of weather data and counts of
lum survival, and mixed populations on aphids in the field to predict the inci­
epidemiology of southern com leaf dence of yellowing viruses of sugar-beet
blight. Phytopathology 64: 1 68-73 crops in England in relation to the use
92. Sutton, T. B., Jones, A. L. 1 976. Evalu­ of insecticides. Ann. Appl. Bioi. 8 1 :
ation of four spore traps for monitoring 1 81-98
discharge of ascospores of Venturia in­ 103. Webster, R. K., Bolstad, J., Wick, C.
aequalis. Phytopathology 66:453-56 M., Hall, D. H. 1 976. Vertical distribu­
93. Vanderplank, J. E. 1 968. Disease R esis­ tion and survival of Sclerotium oryzae
tance in Plants. New York: Academic. under various tillage methods. Phytopa­
206 pp. thology 66:97-101
94. Vanderplank, J. E. 1967. Epidemiology 104. Wiese, M. V . , Ravenscroft, A. V. 1975.
of fungicidal action. In Fungicides: A n Cephalosporium gramineum popula­
Advanced Treatise, ed. D . C . Torgeson, tions in soil under winter wheat cultiva­
1:63-92. New York: Academic tion. Phytopathology 65: 1 129-33
95. Vanderplank, J. E. 1 963. Plant Diseases: 105. Zadoks, J. C. 1 97 1 . Systems analysis
Epidemics and Control New York: and the dynamics of epidemics.
Academic. 349 pp. Phytopathology 6 1 :600- 10

You might also like