You are on page 1of 14

Received: 28 July 2023 | Accepted: 5 November 2023

DOI: 10.1111/ppa.13839

REVIEW ARTICLE

Weather-based models for forecasting Fusarium head blight


risks in wheat and barley: A review

Taurai T. Matengu1 | Paul R. Bullock1 | Manasah S. Mkhabela1,2 | Francis Zvomuya1 |


Maria A. Henriquez1,3 | E. RoTimi Ojo1,4 | W. G. Dilantha Fernando5

1
Department of Soil Science, University of
Manitoba, Winnipeg, Manitoba, Canada Abstract
2
Climate Resilience Unit, Manitoba Fusarium head blight (FHB) is one of the most devastating crop diseases worldwide,
Agriculture, Winnipeg, Manitoba, Canada
3
significantly reducing the yield and quality of small-cereal crops such as wheat and
Morden Research and Development
Centre, Agriculture and Agri-Food barley when favourable weather conditions exist during anthesis. Additionally, FHB-
Canada, Morden, Manitoba, Canada associated mycotoxins significantly impact global food and feed safety. Controlling
4
Soil and Agriculture Weather
FHB with fungicides applied near anthesis reduces visual FHB symptoms and associ-
Surveillance, Manitoba Agriculture,
Winnipeg, Manitoba, Canada ated mycotoxin production, thereby lowering disease-related costs. However, when
5
Department of Plant Science, University weather conditions are unfavourable for FHB occurrence, fungicide application can
of Manitoba, Winnipeg, Manitoba, Canada
be costly and environmentally undesirable. Thus, fungicides should be used sparingly
Correspondence only when the pathogen is present and weather conditions are favourable. Modelling
Paul R. Bullock, Department of Soil
Science, University of Manitoba, of FHB risk using weather data has grown rapidly in recent decades and plays an es-
Winnipeg, Manitoba, Canada. sential role in integrated crop disease management. In this review, several weather-
Email: paul.bullock@umanitoba.ca
based FHB models are selected and described in detail. The models were developed
Funding information globally for assessing the real-time risk of FHB epidemics in various regions/countries.
Agriculture and Agri-Food Canada;
Alberta Wheat Commission; Brewing Most of these models are site-specific and predict FHB visual observations such as
and Malting Barley Research Institute; the incidence and severity of FHB, Fusarium-damaged kernels (FDK), and also deox-
Canadian Agricultural Partnership;
Integrated Crop Agronomy Cluster; ynivalenol (DON) levels. The review also highlights the limitations of these existing
Manitoba Crop Alliance; Saskatchewan models, including their narrow applicability, low accuracy for high-risk contamination
Wheat Development Commission;
Western Grain Research Foundation; situations, and omissions of certain factors. Also discussed are potential avenues for
Prairie Oat Growers Association improvement and enhanced predictive capabilities including consideration of addi-
tional disease risk factors as well as a broader range of varieties. These predictive
models can assist producers, regulatory agencies, and industry to mitigate potential
food and feed security and safety concerns.

KEYWORDS
barley, deoxynivalenol, Fusarium head blight, weather-based models, wheat

This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium,
provided the original work is properly cited.
© 2023 The Authors. Plant Pathology published by John Wiley & Sons Ltd on behalf of British Society for Plant Pathology.

Plant Pathology. 2023;00:1–14.  wileyonlinelibrary.com/journal/ppa | 1


|

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2 MATENGU et al.

1 | I NTRO D U C TI O N countries. The selection criteria included only published field stud-
ies that offer predictions on FHB occurrence in wheat and barley,
Epidemics of Fusarium head blight (FHB) in small-cereal crops such anchored primarily by weather-driven variables. Additionally, only
as wheat and barley are frequently triggered when weather condi- those models with published model equations were considered.
tions favourable for infection by Fusarium species, predominantly The goal is to review a range of FHB risk model examples to identify
Fusarium graminearum, coincide with flowering and early kernel fill- some common characteristics as well as knowledge gaps. When the
ing (De Wolf et al., 2003; McMullen et al., 2012). Numerous studies model developers did not specify a name for a model, the author's
and surveys have established that temperature and moisture (rain- name(s) is used.
fall and humidity) are the primary determinants of FHB develop-
ment (Del Ponte et al., 2009; Hooker et al., 2002; Shah, De Wolf,
et al., 2019). If FHB is left uncontrolled, wheat and barley heads are 2 | FU SA R I U M H E A D B LI G HT CO NTRO L
blighted, and severe yield reductions occur due to flower abortion,
which reduces the number of kernels formed (Bai & Shaner, 2004). Control of FHB outbreaks is crucial due to the substantial loss in
Grain filling is also impacted, leading to Fusarium-damaged kernels crop yield and quality and the adverse health effects of mycotoxins
(FDK), often light in weight (Góral et al., 2018). These yield losses are on consumers. Besides cereals, Fusarium species infect other crops
exacerbated by the production of mycotoxins such as deoxynivale- grown in western Canada, causing diseases such as root rot in soy-
nol (DON) by the causative pathogen (i.e., Fusarium species), which beans, peas and beans; ear rot in maize; and Fusarium wilt in canola
are toxic to both humans and livestock (Tamburic-Ilincic et al., 2015). (Foroud et al., 2019; Gilbert & Fernando, 2004). As a result, FHB
DON-contaminated grains may be unfit for human consumption or disease management with a single strategy such as crop rotation is
the production of products such as bread, beer and animal feed, challenging (Harris et al., 2016). Therefore, it is essential to combine
and are thus frequently downgraded during marketing (Dahl & multiple control strategies to manage FHB (Bai & Shaner, 2004).
Wilson, 2018). Tillage and crop rotation with non-host crops are suitable agri-
Due to the high cost of FHB damage to a crop, some produc- cultural activities that reduce the likelihood of FHB epidemics by
ers use preventative fungicide applications to protect the crop eliminating wintering plant residues and reducing inoculum sources
from disease without knowing if FHB will cause significant damage (Spolti et al., 2014). While reducing inoculum levels and disease epi-
(Nita, 2013). Under low predicted risk, fungicide overuse can be det- demics with high-tillage levels is significant, it is not enough to con-
rimental to the environment, and the marginal cost-to-revenue ratio trol FHB (Gilbert & Fernando, 2004). The selection of varieties with
of wheat makes it critical to use inputs cost-effectively, including some resistance to FHB also plays a crucial role in managing FHB.
fungicides (Wallhead & Zhu, 2017). Fungicide spray for FHB sup- Several FHB resistance types have been identified, for example,
pression in wheat and barley production can be reduced if the risk Type I (resistance to initial infection) and Type II (resistance to patho-
of a disease epidemic can be accurately predicted. This could result gen spread throughout the spike; Góral et al., 2020; Ha et al., 2016).
in more sustainable and environmentally-friendly wheat and barley Additionally, the application of fungicides during flowering greatly
production. Forecasting models are appropriate for FHB due to the reduces FHB epidemics (Yoshida et al., 2012). The efficacy of chemi-
disease's sporadic manifestation, reliance on weather factors, and cal controls relies on predicting the best possible time for fungicide
the relatively short time window for pathogen sporulation, inoculum application, as their effectiveness is much lower if applied either
dispersal and host infection (De Wolf et al., 2003; Góral et al., 2018). prior to or after anthesis (Bolanos-Carriel et al., 2020).
Researchers worldwide have created, validated and adapted weath-
er-based models to predict the likelihood of the presence and sever-
ity of FHB and DON toxin levels so that informed in-season decisions 3 | D I S E A S E FO R EC A S TI N G M O D E L S
can be made and post-season marketing decisions can be anticipated
(Birr et al., 2019; Del Ponte et al., 2005; Rossi, Giosue, Pattori, & Del FHB forecasting has gained popularity in recent decades to maximize
Vecchio, 2003). These models are primarily based on meteorological fungicide efficacy while also minimizing its cost (Musa et al., 2007).
data such as temperature, relative humidity and precipitation, which FHB is ideal for forecasting due to its dependency on environmental
are sometimes combined with agronomic variables such as crop res- conditions and the limited time period near the anthesis stage for
idues, tillage, crop rotation and crop variety. Some of these models host infection (De Wolf et al., 2003; Rossi, Giosue, Pattori, & Del
are specific to where they are developed, so researchers have tested Vecchio, 2003). FHB disease forecasting, days or weeks before in-
these models in different climatic zones to determine how well they fection or major epidemic, helps producers to respond quickly and
work. After a few modifications, some of these models worked well efficiently by modifying crop management practices to include fun-
in other locations and crop types (Giroux et al., 2016; Schaafsma & gicide application (Shah et al., 2013). FHB forecasting has adopted
Hooker, 2007) and sometimes they did not (Landschoot et al., 2013). different strategies including (i) predicting the risk of disease oc-
This review discusses various weather-based FHB models developed currence based on FHB incidence and severity, (ii) forecasting the
around the world for assessing the real-time risk of FHB epidem- FDK risk, and (iii) forecasting the concentration of DON in harvested
ics, as well as their application and adoption in different regions/ grain. Because FHB is a global threat to cereals, the origins for
|

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
MATENGU et al. 3

disease prediction models have come from different research organ- De Wolf et al. (2003) noted that the models with post-anthesis
izations in many different countries. These groups conduct their re- weather variables had higher accuracy than those with just pre-an-
search locally and the approaches that have been developed for FHB thesis weather variables. Model accuracy (i.e., successful classifica-
disease risk prediction have tended to be regional in nature. Thus, tion of FHB epidemics and non-epidemics) was 84%, 84% and 70%
this review has taken the approach to present the models based on for Models A, B and I, respectively. Further validation of Models A
these geographical distinctions in the types of models to provide a and B with an independent dataset showed an accuracy of 78%,
comprehensive understanding of FHB forecasting strategies. which was perceived to be good (De Wolf et al., 2003). Despite its
lower accuracy, Model I was of interest because it had the potential
to provide predictions early enough to make management decisions
3.1 | Research on FHB prediction models from the before anthesis. De Wolf et al. (2003) suggested that the accuracy
United States of FHB prediction models could be improved using new represen-
tations of precipitation variables, such as incorporating the inten-
The first models for FHB risk assessment in wheat in the United States sity and frequency of rainfall events during anthesis, as opposed to
were logistic regression models developed by De Wolf et al. (2003). solely focusing on the duration of precipitation. Additionally, they
These models used data gathered from four states (Ohio, North suggested that incorporating information about local inoculum
Dakota, Missouri and Kansas) for a total of 50 site-years. Field FHB sources, including the presence and type of crop residues, could be
severity ≥10% (congruent to an FHB index ≥10%) was considered useful predictors of disease at a local level, when coupled with key
an epidemic in these forecasting models. Disease forecasting used weather variables.
independent variables constructed from observed hourly weather The De Wolf models were used in an online FHB risk assessment
data for two time periods, either from 7 days before 50% anthesis or tool (i.e., Fusarium Risk Tool; www.​wheat​scab.​psu.​edu), for wheat
10 days after 50% anthesis. A total of 49 variables were constructed production in 31 states across the United States, despite initially
from (i) the calculation of daily minimums, maximums, averages and being developed using data from just four states. While it is unclear
sums; (ii) the duration that prespecified conditions were met; (iii) the how well these four states represent the other states where the
functional response to temperature and moisture; and (iv) the calcu- models were used, the widespread and successful adoption of these
lation of interaction terms between predictor variables to determine models is an indication of their effectiveness in assessing FHB risk
which were most aligned with FHB severity ≥10%. in wheat production. Nevertheless, it is essential to note that the
Eleven different prediction models were developed and three performance of the models may differ between states due to envi-
were selected for closer scrutiny, namely model A, model B and ronmental and crop factors that are unique to each state. The tool
model I (De Wolf A, B and I, respectively). The De Wolf A model provides risk maps showing the probability of a severe FHB outbreak
used only the duration of temperature between 15°C and 30°C and (FHB ≥10%) based on weather data from multiple weather stations in
relative humidity (RH) ≥90% for 10 days post-anthesis (TRH9010; the states covered by the prediction system (McMullen et al., 2012).
Equation 1), while De Wolf B used the interaction of duration of tem- Users enter data such as flowering date, crop type (winter wheat or
perature between 15°C and 30°C for 7 days pre-anthesis (T15307) spring wheat), and crop variety (very susceptible, susceptible, mod-
and TRH9010 (Equation 2). In contrast, De Wolf I used only pre-an- erately susceptible or moderately resistant) into the tool to get FHB
thesis weather variables including T15307 and the duration of pre- risk at the field scale (McMullen et al., 2012; Shah et al., 2021).
cipitation in the 7 days pre-anthesis (DPPT7; Equation 3). More recently, efforts were made by Shah et al. (2013, 2014)
to understand the role of pre-and post-anthesis weather indicators
(1)
( )
P = 1 ∕ (1 + exp − − 3.3756 + 6.8128TRH9010
on FHB epidemic risk forecasting (i.e., FHB epidemics again defined
as an FHB index ≥10%). They used an enormous dataset collected
(2)
( )
P = 1 ∕ (1 + exp − − 3.7251 + 10.5097INT3
across 15 states in the United States over 27 years and an extensive
number of RH, temperature and rainfall predictors for time periods
P = 1 ∕ (1 + exp − − 8.2175 + 8.4358T15307 + 4.7319DPPT7 (3)
( )
ranging from 15 days pre-anthesis to 15 days post-anthesis. These
where P is the probability (from 0 to 1) of FHB severity ≥10%, TRH9010 models added wheat resistance category (RESIT) as an ordinal vari-
is the duration of temperature between 15°C and 30°C and RH ≥90% able, where 0, 1, 2 and 3 represented very susceptible, susceptible,
for 10 days post-anthesis, T15307 is the duration of temperature be- moderately susceptible and moderately resistant, respectively (Shah
tween 15°C and 30°C for 7 days pre-anthesis, DPPT7 is the duration et al., 2013). The odds of FHB index ≥10% on susceptible, moder-
of precipitation in the 7 days pre-anthesis and INT3 is an interaction ately susceptible and moderately resistant cultivars were 47%, 73%
term between T15307 and TRH9010 (De Wolf et al., 2003). To use and 78% lower, respectively, than the odds of a major FHB epidemic
Equations (1)–(3), variables must first be placed on the same scale as on a very susceptible cultivar. Out of 380 total weather-based pre-
the data used to develop the models. This is done by dividing TRH9010, dictors through the various pre- and post-anthesis periods, their
T15307 or DPPT7 by the observed maximum during the study or 136, selection process identified 21 variables as the most important
168 and 39, respectively (TRH9010/136; T15307/168; DPPT7/39) (De predictors, and these were integrated into 15 separate logistic re-
Wolf et al., 2003). gression models. These models differed in the pre- or post-anthesis
|

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
4 MATENGU et al.

duration of their predictors, with specific RH, temperature or rainfall Bondalapati et al. (2012) evaluated and reported significant pos-
conditions (Shah et al., 2013). All 15 models offered a better sensi- itive correlations between DON concentration and both FHB inci-
tivity-specificity balance than updated versions of the original mod- dence and FHB severity in the northern Great Plains of the United
els developed by De Wolf et al. (2003). The presence of a greater States. Using disease and weather data from six growing seasons
number of weather-based predictors in a model was found to be (51 location-years), they developed and validated predictive models
linked to a reduction in misclassification rates. Rainfall was generally for DON epidemics (DON concentration ≥0.5 ppm) in barley using
not crucial for forecasting the likelihood of epidemics in this study logistic regression and the Weibull function to represent the non-
(Shah et al., 2013). Their winter wheat (Equation 4) and spring wheat linear relationship of DON to temperature and wetness. The logis-
(Equation 5) FHB risk models were as follows: tic regression approach was the same as that utilised by De Wolf
et al. (2003) and Shah et al. (2013, 2014). Bondalapati et al. (2012)
logit(𝜇) = − 1.7954 + 0.0245TH2 (4)
developed and evaluated nine prediction models based on different
logit(𝜇) = − 11.008 − 0.9578RESISTC + 0.1516H1 (5) temperature and wetness duration variables near anthesis (±5 days),
which are indicative of ideal conditions for FHB infection. The tem-
where μ is the probability of a significant FHB epidemic (FHB index perature variables included average temperature (AvgT), average
≥10%), H1 is the mean hourly RH in the 7 days pre-anthesis, and TH2 minimum temperature (AvgMinT) and average maximum tempera-
is the number of hours during which the following two conditions are ture (AvgMaxT). The wetness duration variables encompassed hours
met simultaneously within a given hour: temperature of 9°C–30°C and with RH ≥90% (RH90), duration of continuous hours with RH ≥90%
RH ≥90% in the 7 days pre-anthesis. RESISTC is a categorical variable (drRH90), and weighted hours with RH ≥90% (wRH90). Their Model
with four different levels of disease resistance: very susceptible = 0; 3, which used AvgT and the weighted time (in hours) when the RH
moderately susceptible = 1; moderately resistant = 2; and resistant = 3 ≥90% (wRH90) during the critical period (anthesis ±5 days), was found
(Shah et al., 2013). to have the best overall performance (a combination of accuracy,
The US FHB prediction models described above were devel- sensitivity and specificity) for predicting risk of DON ≥0.5 ppm. The
oped using temperature and moisture-based variables during short model accuracy was 84% using the training dataset and 79% using
periods of ≤15 days preceding and following anthesis (De Wolf an independent validation dataset. Bondalapati et al. (2012) noted
et al., 2003; Shah et al., 2013, 2014). This was based on the idea that that, even though cultivar was not a significant factor when included
FHB prediction should be made in time to provide fungicide applica- in the model, their observations indicated that a model with differ-
tion recommendations during anthesis or no later than 5 days after ent thresholds varying with cultivar resistance might provide more
anthesis to control the disease effectively (De Wolf et al., 2003; accurate predictions. They also commented that their study was lim-
Shah et al., 2013) and corresponding to the period of peak suscepti- ited to six growing seasons and did not necessarily represent all of
bility of wheat to the FHB fungal pathogen (Kazan et al., 2012). the possible weather conditions that could occur in the region and
Shah, De Wolf, et al. (2019) and Shah, Paul, et al. (2019) con- influence FHB disease development or DON accumulation in barley.
ducted a functional data analysis of entire weather time series from Therefore, they concluded that further validation was required.
120 days before to 30 days after anthesis at 865 site-years and com-
pared FHB epidemics (FHB index ≥10%) to FHB non-epidemics (FHB
index <10%). The two classes were compared to determine whether 3.2 | Research on FHB prediction models
the mean functional weather curves differed statistically between from Canada
FHB epidemics and non-epidemics. Although the largest differences
between weather time series for epidemics and non-epidemics were In western Canada, some US models have been adopted for pre-
found close to anthesis, the results indicated that FHB epidemics dicting FHB severity greater than 10%. In Manitoba and Alberta,
and non-epidemics have distinct differences in some weather pa- FHB risk assessment maps were developed based on De Wolf I
rameters, particularly atmospheric moisture variables such as RH, as (Equation 3), whereas in Saskatchewan, FHB risk assessment maps
early as 40 days before anthesis (Shah, De Wolf, et al., 2019; Shah, were originally based on the second-generation model developed
Paul, et al., 2019). Differences in the mean functional curves of daily by Shah et al. (2013) that incorporated agronomic and climatic
average temperature between epidemics and non-epidemics were variables (ACIS, 2021; MARD, 2021; SWDC, 2021). The lack of
relatively small and confined to narrow time windows around anthe- validation research on FHB risk assessment models used in the
sis. A better temperature variable was the daily difference between prairie provinces of western Canada can be partly attributed to
the maximum and minimum observed hourly temperature, as had the time-consuming and costly nature of gathering and analys-
been reported previously by Landschoot et al. (2013) in Belgium. The ing such data. A recent study used a logistic regression approach
findings suggest that earlier detection of FHB epidemics is possible similar to De Wolf et al. (2003) to identify weather variables
but must be captured in predictive models that use weather time most closely related to FHB, FDK and DON epidemics in spring
series for variables that are consistently well-separated between wheat, winter wheat, barley and durum wheat on the prairies
epidemics and non-epidemics for extended time periods (Shah, De (Matengu, 2022). A series of several different risk models were
Wolf, et al., 2019; Shah, Paul, et al., 2019). developed using data collected at plot sites in 2019 and 2020 and
|

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
MATENGU et al. 5

then were independently validated using data collected in more of common FHB models across all prairie provinces (Matengu, 2022)
than 200 producer fields across the prairies. Seven of those risk will improve understanding of FHB epidemics and enhance regional
models, shown below (Equations 6–12), were selected to be used collaboration among researchers and stakeholders to promote effec-
in a prairie-wide FHB risk mapping tool (https://​w ww.​youtu​b e.​ tive FHB management.
com/​watch?​v=​z8Ewf​r w_​j6c&​t= ​31s). Forecasting the FHB index is useful but some infected kernels
are, in some instances, asymptomatic even though they are in-
(6)
{ [ ( )]}
FHBww = 1 ∕ 1 + exp − − 5.1095 + 0.0312RH8014
fected with mycotoxin (Hooker et al., 2002). One limitation of the
FHBsw = predictability of DON epidemics is that visual FHB symptoms ob-
(7)
served during grain filling are not necessarily associated with DON
{ [ ( )]}
1∕ 1 + exp − − 34.5786 + 0.3513RHmx14 + 0.0435T252814
incidence (Cowger et al., 2020; Kelly et al., 2015; Matengu, 2022;
Paul et al., 2005). In Ontario, Canada, a model predicting DON con-
{ [ ( )]}
FHBb = 1 ∕ 1 + exp − − 3.7724 + 0.2146R14 + 0.0495T252814
(8) centration in wheat grain was developed by Hooker et al. (2002)
and consisted of three regression equations. The model was further
(9)
{ [ ( )]}
FHBd = 1 ∕ 1 + exp − − 2.0665 + 0.0326TRH8010
improved through the years into a forecasting tool called DONcast
(Schaafsma & Hooker, 2007). In the model, the grain DON level is de-
(10)
{ [ ( )]}
FDKsw = 1 ∕ 1 + exp − − 31.6372 + 0.4004RH10
pendent on three critical periods. The first critical period is 4–7 days
before heading and calculated by Equation 13:
(11)
{ [ ( )]}
FDKd = 1 ∕ 1 + exp − − 11.9932 + 0.0847RH8010 [ ( )2 ]
DON = exp − 0.30 + 1.84RAINA − 0.43 RAINA − 0.56TMIN − 0.1

(12) (13)
{ [ ( )]}
DONd = 1 ∕ 1 + exp − 24.1039 + 0.3114RH14

where FHBww, FHBsw, FHBb and FHBd are the probability values for where DON concentration (mg/g) is a function of RAINA (the num-
FHB index values ≥5% in winter wheat, spring wheat, barley and durum ber of days with rain >5 mm/day), and TMIN is the number of days of
wheat, respectively; FDKsw is the probability for FDK ≥0.3% in spring temperature < 10°C between 4 and 7 days before heading (Hooker
wheat; FDKd is the probability for FDK ≥2.0% in durum; DONd is the et al., 2002). The second critical period is 7 days before heading to
probability for DON ≥1 ppm in durum; RH8014 is the duration (hours) 10 days after heading and calculated using Equation 14:

[ ( )2 ]
DON = exp − 2.15 + 2.21RAINA − 0.61 RAINA + 0.85RAINB + 0.52RAINC − 0.30TMIN − 1.10TMAX − 0.1 (14)

of RH ≥80% in the 14 days prior to mid-anthesis; RHmax14 is the mean where RAINB is the number of days of rain >3 mm/day in the period
daily maximum RH (%) in the 14 days prior to mid-anthesis; T252814 is 3–6 days after heading, TMIN is the number of days with tempera-
the duration (hours) with 28 ≥ air temperature ≥ 25 in the 14 days prior ture < 10°C between 4 and 7 days before heading, and TMAX is the
to mid-anthesis; R14 is the mean daily rainfall (mm) in the 14 days prior number of days with temperature > 32°C in the period 4 and 7 days
to mid-anthesis; TRH8010 is the duration (hours) with both 30 ≥ air tem- before heading. The third critical period included RAINC (number of
perature ≥ 15 and RH ≥80% in the 10 days prior to mid-anthesis; RH10 days of rain >3 mm/day in the period 7–10 days after heading), and is
is the mean daily RH (%) in the 10 days prior to mid-anthesis; RH8010 is given in Equation 15:
the duration (hours) with RH ≥80% in the 10 days prior to mid-anthesis
(15)
( )
DON = exp − 0.84 + 0.78RAINA + 0.40RAINc − 0.42TMIN − 0.1
and RH14 is the mean RH (%) in the 14 days prior to mid-anthesis. The
model accuracies ranged from 75% to 85% on the training data and Rainfall, high humidity and warm temperatures were favour-
from 70% to 100% on the validation data. It should be noted that the able for disease development during all three periods, and DON
dry weather conditions during the study created a limited number of concentrations in the increased set of weather variables for the
epidemic samples in the validation dataset and the accuracy of the dif- three periods varied marginally. Daily rainfall >5 mm increased the
ferent models is mainly or completely a result of the models' specificity. potential DON concentration in the first critical period. In contrast,
Although the Matengu (2022) study was conducted during a drier pe- DON concentration decreased when the daily minimum tempera-
riod with limited FHB disease pressure, the models developed showed ture was <10°C. In the second critical period, DON levels were re-
higher accuracy than the De Wolf Model I and the Shah et al. (2013) duced with mean temperature >32°C. In both the second and third
models being used in the prairie provinces, when they were tested critical periods, daily rainfall >3 mm and RH > 80% increased DON
using the same weather and FHB data. Very recently, Manitoba has concentration in the grain, while temperature < 10°C reduced DON
adopted the spring wheat FHB risk model (Equation 7) as its new in- concentration (Schaafsma & Hooker, 2007).
dicator of FHB risk in the province commencing in the spring of 2023. The model developed by Hooker et al. (2002) explained 73% of
Current research is underway to build a larger dataset of weather and the variability in DON concentration across 5 years and 399 sam-
FHB disease observations on the prairies to improve the accuracy of ples. Prediction of DON concentration lower than 1 mg/L showed
the models and delineate differences in FHB response between culti- a high accuracy of 89%. The Hooker model was further enhanced
vars with varying genetic resistance to the disease. The development by incorporating FHB observations into the dataset and variables of
|

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
6 MATENGU et al.

crop history, tillage and host susceptibility to render it field-specific. simplest models had the best performance despite being developed
The Hooker model was further developed by Weather Innovations in a different geographical region. The De Wolf I model never pro-
(WIN) Consultancy LP into the world's first commercialized DON vided reliable predictions and had low overall accuracy, sensitivity
forecasting system, (i.e., DONcast), accessible in different parts of and specificity. Giroux et al. (2016) commented that the De Wolf I
the world (Schaafsma & Hooker, 2007). DONcast is reasonably sta- model and the Molineros model both use only pre-anthesis weather
ble across the years, continents and cropping systems due to the conditions, which may explain their low performance.
robustness of the data used for its development. For example, one
study found DONcast accuracy of 80%–85% and explained 72% of
DON variation in over 1000 field samples from four countries using 3.3 | Research on FHB prediction models
a 1 mg/L threshold level (Schaafsma & Hooker, 2007). from Argentina
The DONcast model considers a wide range of agronomic fac-
tors, such as varietal susceptibility to disease and the amount of A site and year-specific empirical model to predict the incidence
crop residue on the soil surface. However, it was difficult to get ac- of FHB (predictive index or PI%) was developed by Moschini and
curate DON forecasts and fungicide treatment recommendations Fortugno (1996) for the cooler Pergamino wheat region in Argentina
due to over-generalized agronomic factors and regional variability, using cultivars of different FHB susceptibility, temperature and pre-
which led to low system adoption (Pitblado et al., 2007). Thus, the cipitation as predictor variables. Equations 16 and 17 were used to
DONcast model was refined further by including other meteorologi- predict head blight incidence (PI).
cal variables such as leaf wetness and agronomic factors such as crop
PI( % ) = 20.37 + 8.63 × NP2 − 0.49 × DD926 (16)
history, tillage and crop FHB resistance to make it more field-spe-
cific and improve its forecasting accuracy (Pitblado et al., 2007;
PI( % ) = 18.34 + 4.12 × NP12 − 0.45 × DD1026 (17)
Schaafsma & Hooker, 2007). As a result, the DONcast has been
renamed site-specific DONcast (ssDONcast) (Pitblado et al., 2007; where NP2 is the number of 2-day periods of precipitation ≥0.2 mm,
Schaafsma & Hooker, 2007). and RH > 81% and ≥ 78% for the first and second day, respectively.
A study in Quebec, Canada, evaluated nine different FHB models NP12 is the total NP2 periods plus the number of days with precipi-
for their ability to forecast FHB index, FDK and DON under local tation ≥0.2 mm and average RH > 83%. Maximum and minimum tem-
conditions (Giroux et al., 2016). The models that were tested included peratures were used to calculate degree days between 9 and 26°C.
Hooker et al. (2002), DONcast (Schaafsma & Hooker, 2007), De The observations start 8 days before heading (Zadoks stage 51) and
Wolf A, B and I, Molineros (2007), Moschini et al. (2001), and the two extend up to 530-degree days accumulation. In the equations, DD926
indices developed by Rossi, Giosue, Pattori, and Del Vecchio (2003). and DD1026 represent 926 and 1026 accumulated extreme tempera-
All of these FHB models, except Molineros (2007), are described in tures, respectively, which are calculated using Equations 18 and 19:
separate sections of this paper. Molineros (2007) developed models ∑ [(
(18)
) ( )]
DD926 = Tmax − 26 + 9 − Tmin
based on the De Wolf et al. (2003) models but using only average
RH (%) for the 7 days before flowering and integrating four levels of ∑ [(
(19)
) ( )]
DD1026 = Tmax − 26 + 10 − Tmin
cultivar susceptibility. The De Wolf A, B, I, Molineros, and Moschini
models are empirical and make predictions using polynomial equa-
tions. The Hooker et al. (2002), DONcast and Rossi models are more If the maximum daily temperature is ≤26°C, the accumulation of
complex with FHB risk determined using forecasts at different plant Tmax − 26 is set to zero, and if the minimum daily temperature is ≥9°C,
development stages. the 9 − Tmin is set to zero (Equation 18). Similarly, if the minimum
During the Giroux et al. (2016) study period, there was insuffi- daily temperature is ≥10°C, 10 − Tmin is set to zero (Equation 19).
cient disease pressure to create enough cases with FHB index ≥10%, Therefore, DD values accumulate only on days with extreme high
FHB index ≥5% and DON ≥2 ppm for a statistically robust receiver and low temperatures.
operating characteristic (ROC) analysis. However, the models were Predicted disease severity is estimated using the calculated PI
evaluated using ROC analysis for their accuracy to predict FHB ep- values in Equations (20) and (21).
idemics defined by FDK ≥1.5%, FDK ≥1.0% and DON ≥1 ppm be-
PSms = − 1.010225 + exp(0.137449 × PI) (20)
cause there were a sufficient number of epidemic cases for robust
analysis. The DON ≥1 ppm disease indicator best differentiated the
PSs = − 1.02573985 + exp(0.27209 × PI) (21)
differences in the performance of the nine models. The De Wolf A,
B and Moschini models all displayed a good combination of accu- where PSms is predicted head blight severity (0 to 1) for moder-
racy, sensitivity, and specificity for DON ≥1 ppm, even though these ately susceptible wheat cultivars, PSs is predicted head blight se-
models were not designed to predict DON levels directly but rather verity (0 to 1) for susceptible wheat cultivars and PI is calculated
field FHB index. The Molineros model underperformed with re- in Equations (16) and (17). Equations (20) and (21) were developed
gard to specificity and the Hooker et al. (2002), DONcast and Rossi from 64 and 31 data points, respectively with R 2 values of 0.947 and
models underperformed in regards to sensitivity. Thus, some of the 0.907, respectively.
|

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
MATENGU et al. 7

When the equations above were used in warmer wheat regions Y95 = dummy variable for 1995, equal to zero (for the years 1994, 1996
of Argentina, there was a tendency for underestimation at high dis- and 1997) and 1 (for 1995); RH80 = number of hours with RH > 80%;
ease levels. The model was adapted to these warmer regions in more and Rtot_P = total rainfall (mm) on a preceding day (Rossi et al., 2002).
northern Argentina (Moschini et al., 2001). The maximum daily tem- Disease infection frequency (INF) was based on leaf wetness du-
perature threshold in Equations (18) and (19) was changed from 26°C ration and temperature following inoculation and considering wheat
to 30°C and the number of degree days in the critical period was growth stages (GS) (Rossi, Giosue, Pattori, & Del Vecchio, 2003). For
increased from 530 to 550. The revised model offered predictions G. zeae, INF was estimated using the following equation:
closer to the observed infection levels (Moschini et al., 2001). Unlike
INF = − 0.099 − 0.363t + 0.07808 (T × t) − 0.00591 T 2 × t
( )
logistic regression models, this Argentinian model does not have an ( 3 ) ( 4 ) (25)
+ 0.000199 T × t − 0.0000024 T × t
action threshold (i.e., a specific level that triggers fungicide appli-
cation). It directly predicts the level of FHB incidence and severity where T = temperature (°C) and t = incubation time (hours).
based on air temperature and precipitation during the critical time Finally, the invasion rate of mycelium into head tissues (INV) was
period for infection (Moschini et al., 2001). related to temperature. For G. zeae, INV was estimated using the
following equation:

3.4 | Research on FHB prediction models from Italy


)]1.35
(26)
[ (
INV = 5.33 × Teq 1.55 × 1 − Teq

Rossi et al. (2002); Rossi, Giosue, Pattori, and Del Vecchio (2003); where Teq = T − Tmin ∕ Tmax − Tmin , Tmin = 0 and Tmax = 38.
( ) ( )

Rossi, Giosue, and Delogu (2003) developed an epidemiological Two risk indices are given by the Rossi model and are calculated
model in Italy, in which weather variables and information on wheat as below:
growth stages were used to predict the risk of FHB and DON. Model
FHB _ risk = ΣSPO × DIS × INF × GS (27)
development was based on four problematic FHB pathogens in Italy
(Fusarium culmorum, Fusarium graminearum, Fusarium avenaceum
TOX _ risk = ΣINF × GS × INV (28)
and Microdochium nivalis). The model pivots on equations related
to the phases in the epidemiological cycle of FHB in wheat. In the The risk of FHB infection (FHB_risk) is used for the four Fusarium
first equation, the amount of inoculum produced, or the sporulation species and the risk of mycotoxin development (TOX_risk) for F.
rate (SPO), was determined using four equations (one for each fun- culmorum and F. graminearum (Rossi, Giosue, & Delogu, 2003). The
gal species) under controlled incubation conditions (Rossi, Giosue, indices are computed daily and accumulate throughout the grow-
Pattori, & Del Vecchio, 2003). In Gibberella zeae, SPO was estimated ing season. Data from various winter wheat crops were used to
using the following equation: validate the model. A comparison of actual infection and mycotoxin
against predicted data produced satisfactory results (R 2 > 0.8) (Rossi,
)]0.24 [
Giosue, & Delogu, 2003).
[
8.59
∕ 1 + exp(5.52 − 0.51t) (22)
( ]
SPO = 25.98Teq × 1 − Teq

where Teq = equivalent of temperature calculated as (T − Tmin)/


(Tmax − Tmin); T = temperature (Tmin = 5°C and Tmax = 35°C) and t = incuba- 3.5 | FHB prediction model research from Brazil
tion time (days).
Spore dispersal rate (DIS) was determined using precipitation, A process-based FHB simulation model was developed in Brazil, ac-
which is the main factor influencing the rate of spore dispersal. Two counting for host, environment and inoculum dynamics during in-
regression equations were developed, one adjusted to rainy days fection (Del Ponte et al., 2005). Firstly, the host factor calculates
(rain >0.2 mm) and the other for non-rainy days (rain <0.2 mm). The the groups of heads that emerged (cohort) on the same day (HNG)
regression equations also included rainfall intensity, daily average (Equation 29). The daily rate in the cohort of heads as a function of
temperature, and duration of hours with RH > 80% as shown below temperature (ANText) was then calculated (Equation 30).
(Rossi et al., 2002):
HNG = 1 − exp − 0.0127t 2.4352 (29)
( )

DIS1 = − 839.7 + 410.3W + 4.08T 2 + 115.45RintMax_P − 455.9Y95 (23)


ANText = 1 − exp at b (30)
( )

DIS2 = − 682.3 + 45.68W + 21.5T 2 + RH80 + 107.0Rtot_P (24)


where t = 1 day, a = 0.225–0.029T + 0.0009T2 and
3
where DIS1 and DIS2 are conidia numbers (m of air per day) estimated b = −5.773 + 0.966T − 0.0278T2, and T is daily mean temperature
in days with and without rainfall, respectively; W is an empirical weight (°C). Equation (29) assumes that every part of the cohort has its
assigned each day in a sequence of n successive rainy days, as fol- first part extruded 3 days later. An empirical rule was determined
lows: first rainy day = 1.1, second rainy day = 2.5, third rainy day = 1.2, when anthers were attached to wheat spikelets before falling on
fourth or later rainy day = 0.8; T is the average air temperature (°C); the ground (anther longevity). A rule was established based on em-
RintMax_P = maximum intensity of rain (mm/h) in the preceding day; pirical observations reporting prolongation of flowering during a
|

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
8 MATENGU et al.

series of cloudy days, indicating that flowers remained attached for where ANT is the daily mean proportion of anthers during a 2-day
longer. The rule suggested that flower longevity is a minimum of infection event (IE), ranging from 0 to 1; ST = mean daily proportion
2 days and, if the daily solar radiation on the second or next day is of susceptible tissue during IE; INF = infection frequency at the sec-
<10 MJ m−2 day−1, the anther remains attached for an extra day up ond day of IE and GZ = mean spore cloud density during IE. The accu-
to a maximum of 5 days. Therefore, the proportion of anthers pres- mulated infection index (% GIB) is the summation of partial infection
ent in 1 day (ANT) results from the sum of extruded and attached indices by the four different models along the susceptible period
anthers in each cohort of heads subtracted from anthers removed (Equation 37):
from the cohorts (Del Ponte et al., 2005).
GIB % = Σ(GIB × 100) (37)
Coefficients for determining the proportion of susceptible tis-
sue (ST) were developed using ANT and coefficients for post-peak The models were evaluated based on FHB incidence, FHB sever-
anthesis infections. ST is ANT until ANT reaches the peak and de- ity and FDK. Because the models estimate an infection index rather
creases to 0.25. ST = 0.25 for the next 7 days after anthesis (ANT than a disease level, regression was used to validate each model
<0.01), while ST = 0.10 for the next 8–14 days after anthesis. Such using the coefficient of determination to verify the adequacy of the
guidelines were developed to compensate for potential late infec- index to explain the observations of the disease level (Del Ponte
tions from post-peak flowering to kernel filling levels (Del Ponte et al., 2005). The coefficient of determination for the regression
et al., 2005). analysis between predicted and actual infection ranged from 0.73 to
Secondly, the inoculum factor was derived from the night and 0.93, 0.43 to 0.69, and 0.14 to 0.37 for FHB severity, incidence and
daytime inoculum observations. Spore cloud relative density was FDK, respectively.
estimated by adjusting a linear equation to the relative density of
colony-forming units observed during the night using Equation (31).
3.6 | Research on FHB prediction models
GZ = ( − 0.6306 + 0.0152RH + 0.176CRD)2 (31)
from Germany
where RH = daily mean relative humidity (%) and CRD is a variable for
the rainy-day position (>0.3 mm) in consecutive rainy days (for four Birr et al. (2019) developed three multiple regression models to
consecutive days: CRD = 1, 2, 2.5 or 0.3 for each subsequent day); and predict concentrations of DON and zearalenone (ZEA) mycotoxin
GZ is a fraction (0 < GZ <1), which adjusts the daily index of infections in wheat grain at harvest in a maize-free crop rotation in northern
by accounting for lower or higher inoculum pressure in the course of Germany. The model for the highly susceptible cultivar (Model 1)
an infection event. was based on data collected from eight farmers' fields from 2008
Thirdly, daily rainfall and mean RH combinations were used to de- to 2017. The moderately to highly susceptible cultivar (Model 2) and
termine head wetness duration between 30 and 48 h. Within 2 days, the low to moderately susceptible cultivar (Model 3) were based on
infection events were recorded, and infection happens when precip- data from 2012 to 2016. The models used cumulative precipitation
itation (PREC) > 0.3 mm occurs in both days with average RH > 80% and average temperature covariates and the interaction term of
for the 2 days or when PREC >0.3 mm with mean RH ≥80% followed precipitation and temperature during wheat flowering. In collabo-
by a non-rainy day with mean RH ≥85%. An exponential equation to ration with the Schleswig-Holstein Plant Protection Service (SCA),
calculate F. graminearum infection frequency (INF) under tempera- the three models were adapted to farms using representative local
ture effect (10°C to 30°C) for 48 h of head wetness was developed weather stations for weather-based DON and ZEA predictions as
as follows: part of the regional monitoring of leaf pathogens and FHB in wheat
in northern Germany.
INF = 0.001029 exp(0.1957T) (32)
Model 1: highly susceptible (Ritmo).
In this case, T = average mean daily temperature in the 2-day win-
DON = 646 − 4.50P − 31.94T + 2.70PT (38)
dow of the infection event.
Finally, four models calculating the daily infection risk index (GIB)
ZEA = − 40 + 3.86P + 2.41T + 0.18PT (39)
were developed. The daily GIB is then added and multiplied by 100
to get the accumulated infection index (% GIB). The GIBs are calcu- Model 2: moderately to highly susceptible (Inspiration).
lated using Equations (33)–(36).
DON = 481 − 18.42P − 26.37T + 2.87PT (40)
GIB1 = ANT × INF (33)
ZEA = − 75 + 2.17P + 5.34T + 0.14PT (41)
GIB2 = ANT × INF × GZ (34)
Model 3: low to moderately susceptible (Dekan).

GIB3 = ST × INF (35) DON = 305 − 24.09P − 18.74T + 2.73PT (42)

GIB4 = ST × INF × GZ (36) ZEA = − 39 + 1.18P + 2.87T + 0.11PT (43)


|

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
MATENGU et al. 9

where P is the cumulative precipitation (mm) during wheat flowering for the entire pre- and post-heading time blocks of 24 days each)
(GS61 to GS69), T is the average temperature (°C) during wheat flow- (Franz et al., 2009). The heading date was estimated by linear re-
ering (GS61 to GS69), PT is the interaction term of precipitation and gression between the observed heading date and the pre-heading
temperature. Model evaluation with independent data not used in the temperature sum (Tsum) when not recorded. The first model focused
development of the models but from similar sampling locations and on average climatic variables in a 24-day pre-heading and a 24-day
years produced satisfactory results (Birr et al., 2019). The coefficients post-heading period. The second model focused on average climatic
of determination for Models 1, 2 and 3 were 0.89, 0.91 and 0.86, re- variables in eight-time blocks of 6 days. Models 1 and 2 explained
spectively, indicating that the variation in the predicted DON values ac- 59% and 56% of DON variation across all fields, respectively, indi-
counted for a substantial proportion of the observed variation in DON cating good predictive performance. Climate variables accounted for
values. In 95.2% of the cases, Model 1 correctly predicted whether 30% of the overall variance in Model 1.
DON and ZEA concentrations were lower or higher than the European
limit of 1250 μg DON/kg and 100 μg ZEA/kg for both mycotoxins. The
accuracy of Models 2 and 3 were 85.7% (14.3% false negative) and 3.7.1 | Model 1
100%, respectively, for DON predictions when the 2017 dataset was
used (Birr et al., 2019). Models 2 and 3 accurately estimated in 100% DON (𝜇g∕kg) = 3.771 + 0.209T (post − heading) + 0.005Pmm (pre − heading)

and 85.7% (14.3% false positive) of cases, ZEA concentrations below + 0.004Pmm (post − heading) − 0.002Th25 (pre − heading)
or above the maximum level of 100 μg ZEA/kg, respectively, when the − 0.005Th25 (post − heading) + 0.001RHh90 (post − heading)
2017 validation dataset was used (Birr et al., 2019). − 0.031HD + REG + RES + SPRAY
In contrast to other published models (Franz et al., 2009; Hooker (44)
et al., 2002; van der Fels-Klerx et al., 2010), all three models of where T (post-heading) is the average hourly post-heading tempera-
Birr et al. (2019) produced reliable DON and ZEA concentration ture, Pmm (pre-heading) and Pmm (post-heading) are the total hourly
estimates in wheat grain without including post-anthesis weather pre- and post-heading precipitation (mm), respectively, Th25 (pre-head-
conditions. Therefore, the models are used not only as a tool for ing) and Th25 (post-heading) are the number of hours with a tempera-
identifying years, regions, or fields under risk, but also to assist ture greater than 25°C and RHh90 is the number of post-heading hours
producers in determining whether to make a fungicide application with RH > 90%. Non-climatic variables included HD (as factorial),
at flowering based on weather data from local weather stations in which are the heading dates; REG is the wheat-producing region with
order to reduce the risk of food and feed contamination with DON coefficients: C REG (reference level); −0.366SW REG and 0.3442NE
and ZEA (Birr et al., 2019). The models use only temperature and REG. RES is the wheat resistance category with coefficients: RES 5
precipitation as input variables, which may not capture all factors (reference level); −0.415RES 5.5; 0.057RES 6; 0.375RES 6.5; 0.482RES
that affect disease development such as tillage practices and crop 7; 0.336RES 7.5; 0.525RES 8.5). SPRAY (factorial) is a fungicide ap-
rotation. Also, the models are based on linear relationships between plication with coefficients: No (reference level); −0.192Yes (Franz
weather variables and DON and ZEA concentrations, which may not et al., 2009).
hold true in all cases, affecting model accuracy. These models are
still relatively new, and further refinement may be needed to im-
prove their accuracy and usability. 3.7.2 | Model 2

DON (𝜇g∕kg) = 4.427 + 0.1320T (8) + 0.025RH (4) − 0.018Th25 (8)


3.7 | Research on FHB prediction models from the − 0.034HD + REG + RES + SPRAY
Netherlands (45)
where HD is heading date and T (8) is the average hourly tempera-
In the Netherlands, the application of the DONcast regression equa- ture in the eighth of eight 6-day time blocks bracketing HD (i.e. 18
tions developed by Hooker et al. (2002) to Dutch data resulted in –24 days after heading), RH (4) is the average hourly relative humidity
poor quantitative predictive performance (Franz et al., 2009). The in the fourth of eight 6-day time blocks bracketing HD (i.e., 0–6 days
poor performance of the DONcast model was attributed to the fact before heading), and Th25 (8) is the number of hours with a T greater
that predictive models for DON levels in wheat may not be univer- than 25°C in the eighth of eight 6-day time blocks bracketing HD
sally applicable. Thus, two multiple regression models specific to the (i.e., 18–24 days after heading). REG is the region where the wheat
Dutch conditions were developed to predict DON concentrations in is produced (factorial) with coefficients as C REG, reference level;
wheat grain in Dutch-specific situations (Franz et al., 2009). These −0.385SW REG; 0.259NE REG, RES is wheat resistance (factorial) with
models use pre- and post-heading weather data, cultivar resistance, coefficients: −0.441RES 5.5; −0.037RES 6; −0.319RES 6.5; −0.508RES
region (REG) classified into south-west (SW), central (C) and north- 7; −0.270RES 7.5; −0.618RES 8.5 and SPRAY (factorial) with coeffi-
east (NE), fungicide use (SPRAY) classified into Yes or No, and head- cients No, reference level; −0.211Yes (Franz et al., 2009).
ing dates (HD) which include eight blocks of 6 days around the HD Both models predict concentrations of 10 log(x + 1) in DON
(numbered 1 to 8 [as a suffix] to the weather variables as well as (μg/kg). Models 1 and 2 correctly predicted whether the DON
10

TA B L E 1 Risk prediction models developed worldwide for wheat and barley Fusarium head blight (FHB) and deoxynivalenol (DON).
|

Input variables for days before [−]/


Country Authors Modelling technique after [+] anthesis Response variables Model performance Limitations

USA De Wolf et al. (2003) Logistic regression Hours of temp. 15°C–30°C and FHB index (spring/winter Accuracy 70%–84% Low accuracy, does not consider
RH ≥90% [+10], hours of temp. wheat) (training data) agronomic practices
15°C–30°C [−7], hours of rainfall
[−7]
Shah et al. (2013, Artificial neural networks FHB genetic resistance mean RH [−7], FHB index (spring/winter Misclassification rate Low accuracy, does not consider
2014) and decision trees within hours of temp. 9°C–30°C and RH wheat) 0.26 (training data), agronomic practices
logistic regression ≥90% [−7] 0.34 (test data)
Bondalapati Logistic regression Hours of RH ≥90% [±5] FHB index, DON (barley) 84% (training data) 79% Those above plus site-specific
et al. (2012) (test data)
Canada Matengu (2022) Logistic regression Hours of RH ≥80% [−14], mean daily FHB index (winter/spring 75%–85% accuracy Low accuracy, limited dataset
max. RH% [−14], hours of temp. wheat, durum, barley), (training data) with low number of epidemic
25°C–28°C [−14], mean daily FDK (spring wheat, 70%–100% accuracy samples, does not consider
rainfall (mm) [−14], hours of temp. durum), DON (durum) (test data) agronomic practices
15°C–30°C and RH ≥80% [−10],
mean daily RH% [−10], hours of RH
≥80% [−10], mean daily RH% [−14]
Hooker et al. (2002) Multiple regression Days with rainfall >5 mm [−4 to −7 DON (wheat) R2 = 0.73 Poor performance reported in
heading], days with min. temp. some region-specific studies
<10°C [−4 to −7 heading], days
with rainfall >3 mm [+3 to +6
heading], days with max. temp.
>32°C [−4 to −7 heading], days
with rainfall >3 mm [+7 to +10
heading]
Argentina Moschini and Degree-day model for the Number of 2-day periods of rainfall FHB incidence and severity R2 = 0.947 (moderately Site and year specific, limited to
Fortugno (1996), critical infection period, ≥0.2 mm and RH > 81% (1st day) (wheat) susceptible moderately susceptible and
Moschini regression models for and RH ≥78% (2nd day), max. and varieties), R2 = 0.907 susceptible varieties
et al. (2001) fusarium incidence and min. temp. (susceptible varieties)
severity
Italy Rossi et al. (2002, Regression models for Spore incubation time, temp., total FHB risk of infection, and R2 = 0.51 (FHB) R2 = 0.90 Low accuracy for high FHB and
2003, 2003) epidemiological stages of rainfall, rainy days (≥0.2 mm), non- DON (spring and winter (DON) DON cases
sporulation (rate, spore rainy days, hours of RH >80%, leaf wheat)
dispersal), infection wetness
frequency, disease
invasion rate
Brazil Del Ponte Second order polynomial Daily mean temp., daily mean RH, FHB infection index, related For the GIB4% model: Applicable only to the specific
et al. (2005) model for daily heading consecutive rainy days to FHB severity, incidence R2 = 0.93 (FHB region studied
and inoculum production, and FDK severity) R2 = 0.69
non-linear regression for (FHB incidence),
infection frequency R2 = 0.37 (FDK)
MATENGU et al.

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
|

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
MATENGU et al. 11

concentration was lower or higher than the maximum level of

tillage practice on DON and


ZEA not directly included in

regions, requires validation


1250 μg/kg in 92% and 88% of the cases, respectively. The estimated

Developed for Dutch-specific


Effects of previous crop and

for use in other regions


DON amount increased in both models with increasing average tem-
perature, precipitation and RH but decreased with an increased
number of hours at temperatures above 25°C (Franz et al., 2009).

the models
There was an apparent regional effect on DON concentrations,
Limitations

with considerably greater levels when progressing from the south-


west to the north-east regions of the Netherlands that climatic and
non-climatic variables could not explain (Franz et al., 2009). Region
0.84 (model 2), 0.89
effects were thought to be caused by differences in infection pres-
0.86 (model 3) ZEA
R2 = 0.91 (model 1),

sure in different parts of the country because of how much wheat is

model 2: R2 = 0.56,
DON R = 0.89 (model
1), 0.91 (model 2),

sensitivity = 0.63;

sensitivity = 0.59
Model performance

grown. Wheat-after-wheat rotations are more typical in the north-


Model 1: R2 = 0.59,

east, with more intensive wheat cultivation. Another possible ex-


(model 3)

planation for the region effect were differences in temperature


2

sensitivity between populations of Fusarium species in different loca-


tions (Franz et al., 2009). Northern populations may have adapted to
lower average temperatures, resulting in increased growth, and divid-
zearalenone (ZEA) (winter

ing regional impact into underlying variables could greatly increase


model accuracy, making prediction models less site-specific and gen-
Deoxynivalenol (DON),

DON (winter wheat)

erally applicable to producers, industry and control authorities (Franz


Response variables

et al., 2009).
wheat)

4 | CO N C LU S I O N
total hourly pre- and post-heading

resistance, sprayed or not sprayed


Cumulative rainfall and average temp.

This paper reviews and summarizes several weather-based mod-


Average hourly post-heading temp.,

post-heading RH >90%, cultivar


Input variables for days before [−]/

rainfall, hours of pre- and post-


during wheat flowering period

heading temp >25°C, hours of

els for FHB disease forecasting in cereal crops that have been de-
veloped and used by a number of researchers in the past couple
of decades (Table 1). From the models reviewed in this study, it is
evident that there is a limited number that directly predict FDK
and DON levels. Moreover, the majority of these models focus on
after [+] anthesis

spring wheat, with only a few addressing barley and durum wheat.
This highlights a significant research gap and emphasizes the need
for the development of more comprehensive and diverse models
capable of accurately predicting FDK and DON levels in a variety
of cereal crops, including durum wheat and barley. While many of
Multiple regression models

Multiple regression models

the weather-based FHB models examined in this study, like the De


Wolf models in the United States, employ regression modelling,
Modelling technique

there are much more complex models, such as those developed


by Rossi for Italy and the Del Ponte model in Brazil, that simulate
the entire dynamic cycle of the FHB pathogen. The quest for iden-
tifying a crucial weather period conducive to FHB epidemics is
common in all models reviewed in this paper. The timeframe for
prediction of the FHB and the probability of outbreak varies from
Franz et al. (2009)

120 days before to 14 days after the commencement of anthesis.


Birr et al. (2019)

Most models produce more accurate predictions using weather pa-


rameters for periods within 14 days before and after anthesis. The
TA B L E 1 (Continued)

Authors

evidence for validity of these models is how well they link theoreti-
cal, scientific and small-scale knowledge to large-scale knowledge
of the disease, and this is assessed by different measures of ac-
Netherlands

curacy (R 2 and accuracy sensu stricto) (Hollingsworth et al., 2006).


Germany
Country

Parsimony, the principle of modelling that favours simplicity rather


than complexity, is preferable (Prandini et al., 2009; Rossi, Giosue,
|

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
12 MATENGU et al.

Pattori, & Del Vecchio, 2003). Of the models reviewed in this C O N FL I C T O F I N T E R E S T S TAT E M E N T
study, Birr et al. (2019) and De Wolf et al. (2003) are examples of The authors declare that there is no conflict of interest.
the most parsimonious models, while Del Ponte et al. (2005) and
Rossi et al. (2002), Rossi, Giosue, Pattori, and Del Vecchio (2003), DATA AVA I L A B I L I T Y S TAT E M E N T
Rossi, Giosue, and Delogu (2003) are considered less parsimonious Data sharing is not applicable to this article as no new data were cre-
models, as they include more complex epidemiological factors that ated or analysed in this study.
account for disease development. Models must be as accurate as
possible while using the fewest possible easily measured variables. ORCID
While complex models can be more accurate and portable than Taurai T. Matengu https://orcid.org/0009-0003-3883-190X
empirical models based on statistical relationships, complexity can W. G. Dilantha Fernando https://orcid.
reduce model usability, adaptability and applicability. The long-term org/0000-0002-2839-1539
accuracy of models may be impacted by effects of climate change
on FHB species evolution, which could lead to epidemic variation REFERENCES
and weather patterns outside those captured in the data used to ACIS. (2021) Alberta Climate Service -Fusarium Head Blight Map.
develop the reviewed models. In the future, it is likely that continu- Available from: https://​agric​ulture.​alber​t a.​c a/​acis/m#​!fusarium.
[Accessed 16th November 2021].
ous data collection for model development, adaptation and ongo-
Bai, G. & Shaner, G. (2004) Management and resistance in wheat and
ing validation will be essential (Vaughan et al., 2016). The recently barley to Fusarium head blight. Annual Review of Phytopathology, 42,
developed logistic regression risk models for FHB index (FHBi), FDK 135–161.
and DON in spring wheat, winter wheat, durum wheat and barley Birr, T., Verreet, J.-A. & Klink, H. (2019) Prediction of deoxynivalenol and
zearalenone in winter wheat grain in a maize-free crop rotation
in western Canada (Matengu, 2022) are expected to be continu-
based on cultivar susceptibility and meteorological factors. Journal
ously updated and validated in the future. The correlation between of Plant Diseases and Protection, 126, 13–27.
these FHB damage indicators is variable, and the risk of infection Bolanos-Carriel, C., Wegulo, S.N., Baenziger, P.S., Funnell-Harris, D.
varies by crop type. These models power an online viewer that pro- & Hallen-Adams, H.E. (2020) Effects of fungicide chemical class,
fungicide application timing, and environment on Fusarium head
cesses near real-time weather information to generate FHBi, FDK
blight in winter wheat. European Journal of Plant Pathology, 158,
and DON estimates. Successful management of FHB will require 667–679.
not just the development of FHB risk models but also collaboration Bondalapati, K.D., Stein, J.M., Neate, S.M., Halley, S.H., Osborne, L.E. &
with local agriculture offices and agronomists as well as extension Hollingsworth, C.R. (2012) Development of weather-based predic-
staff to educate producers about the disease and the best use of tive models for Fusarium head blight and deoxynivalenol accumula-
tion for spring malting barley. Plant Disease, 96, 673–680.
integrated management practices, including the risk assessments
Cowger, C., Ward, T.J., Nilsson, K., Arellano, C., McCormick, S.P. &
that are provided by the models. Disease risk models should be in- Busman, M. (2020) Regional and field-specific differences in
tegrated into disease management decisions using a multifaceted Fusarium species and mycotoxins associated with blighted North
approach. Web-based and smartphone app-based decision aid tools Carolina wheat. International Journal of Food Microbiology, 323,
108594.
enable widespread access to risk information. However, only a small
Dahl, B. & Wilson, W.W. (2018) Risk premiums due to Fusarium head
number of the models examined here, including the De Wolf mod- blight (FHB) in wheat and barley. Agricultural Systems, 162, 145–153.
els, the Shah models and DONcast, have been used to create a suc- De Wolf, E.D., Madden, L.V. & Lipps, P.E. (2003) Risk assessment models
cessful commercial or public online tool for assessing Fusarium risk for wheat Fusarium head blight epidemics based on within-season
weather data. Phytopathology, 93, 428–435.
(Table S1). Prediction of FHB days before anthesis is an essential
Del Ponte, E.M., Fernandes, J.M.C. & Pavan, W. (2005) A risk infection
element of any FHB model if it is to be used as a tool for produc- simulation model for Fusarium head blight of wheat. Fitopatologia
ers to decide on fungicide spraying. Models that use post-anthesis Brasileira, 30, 634–642.
weather conditions to evaluate the risk of FHB will not provide the Del Ponte, E.M., Fernandes, J.M.C., Pavan, W. & Baethgen, W.E. (2009)
A model-based assessment of the impacts of climate variability on
assessment in time for producers to intervene, although they can
Fusarium head blight seasonal risk in southern Brazil: modeling im-
still be helpful for anticipating FHB impacts on grain marketing and pacts of climate variability on Fusarium head blight. Phytopathology,
food systems (De Wolf et al., 2003; Rossi, Giosue, Pattori, & Del 157, 675–681.
Vecchio, 2003). Foroud, N.A., Baines, D., Gagkaeva, T.Y., Thakor, N. & Badea, A. (2019)
Trichothecenes in cereal grains—an update. Toxins, 11, 634.
Franz, E., Booij, K. & Van der Fels-Klerx, I. (2009) Prediction of deoxyni-
AC K N O​W L E D
​ G E ​M E N T S valenol content in Dutch winter wheat. Journal of Food Protection,
We are grateful for the financial assistance provided by the 72, 2170–2177.
Integrated Crop Agronomy Cluster, Western Grain Research Gilbert, J. & Fernando, W.G.D. (2004) Epidemiology and biological con-
Foundation, Agriculture and Agri-Food Canada, Manitoba Crop trol of Gibberella zeae/Fusarium graminearum. Canadian Journal of
Plant Pathology, 26, 464–472.
Alliance, Saskatchewan Wheat Development Commission, Alberta
Giroux, M.-E., Bourgeois, G., Dion, Y., Rioux, S., Pageau, D., Zoghlami, S.
Wheat Commission, Brewing and Malting Barley Research Institute, et al. (2016) Evaluation of forecasting models for Fusarium head
Canadian Agricultural Partnership, and Prairie Oat Growers blight of wheat under growing conditions of Quebec, Canada. Plant
Association. Disease, 100, 1192–1201.
|

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
MATENGU et al. 13

Góral, T., Wiśniewska, H., Ochodzki, P., Nielsen, L., Walentyn-Góral, D. deoxynivalenol accumulation in harvested wheat grain: a me-
& Stępień, Ł. (2018) Relationship between Fusarium head blight, ta-analysis. Phytopathology, 95, 1225–1236.
kernel damage, concentration of Fusarium biomass, and Fusarium Pitblado, R., Hooker, D.C., Nichols, I., Danford, R. & Schaafsma, A.W.
toxins in grain of winter wheat inoculated with Fusarium culmorum. (2007) DONCAST: seven years of predicting DON in wheat on
Toxins, 11, 2. a commercial scale. In: Proceedings of the National Fusarium Head
Góral, T., Wiśniewska, H., Ochodzki, P., Twardawska, A. & Walentyn- Blight Forum. The Westin crown center. Kansas City, Missouri.
Góral, D. (2020) Resistance to fusarium head blight, kernel damage Available from: https://​scabu​sa.​org/​pdfs/​forum​07_​proc_​compl​ete.​
and concentration of Fusarium mycotoxins in grain of winter triti- pdf. [Accessed 6th May 2022]: ASAP Printing.
cale (x Triticosecale Wittmack) lines. Agronomy, 11, 16. Prandini, A., Sigolo, S., Filippi, L., Battilani, P. & Piva, G. (2009) Review of
Ha, X., Koopmann, B. & von Tiedemann, A. (2016) Wheat blast and predictive models for Fusarium head blight and related mycotoxin
Fusarium head blight display contrasting interaction patterns on contamination in wheat. Food and Chemical Toxicology, 47, 927–931.
ears of wheat genotypes differing in resistance. Phytopathology, Rossi, V., Giosue, S. & Delogu, G. (2002) Dynamics of airborne of
106, 270–281. Fusarium macroconidia in wheat fields naturally affected by head
Harris, L.J., Balcerzak, M., Johnston, A., Schneiderman, D. & Ouellet, T. blight. Journal of Plant Pathology, 84, 53–64.
(2016) Host-preferential Fusarium graminearum gene expression Rossi, V., Giosue, S. & Delogu, G. (2003) A model estimating risk for
during infection of wheat, barley, and maize. Fungal Biology, 120, Fusarium mycotoxins in wheat kernels. Aspects of Applied Biology,
111–123. 68, 229–234.
Hollingsworth, C.R., Mewes, J.J., Motteberg, C.D. & Thompson, W.G. Rossi, V., Giosue, S., Pattori, E. & Del Vecchio, A. (2003) A model esti-
(2006) Predictive accuracy of a Fusarium head blight epidemic risk mating the risk of Fusarium head blight on wheat. EPPO Bulletin,
forecasting system deployed in Minnesota. Plant Health Progress, 33, 421–425.
7, 4. Schaafsma, A.W. & Hooker, D.C. (2007) Climatic models to predict occur-
Hooker, D.C., Schaafsma, A.W. & Tamburic-Ilincic, L. (2002) Using rence of Fusarium toxins in wheat and maize. International Journal of
weather variables pre- and post-heading to predict deoxynivalenol Food Microbiology, 119, 116–125.
content in winter wheat. Plant Disease, 86, 611–619. Shah, D.A., De Wolf, E.D., Paul, P.A. & Madden, L.V. (2014) Predicting
Kazan, K., Gardiner, D.M. & Manners, J.M. (2012) On the trail of a cereal Fusarium head blight epidemics with boosted regression trees.
killer: recent advances in Fusarium graminearum pathogenomics and Phytopathology, 104, 702–714.
host resistance. Molecular Plant Pathology, 13, 399–413. Shah, D.A., De Wolf, E.D., Paul, P.A. & Madden, L.V. (2019) Functional
Kelly, A.C., Clear, R.M., O'Donnell, K., McCormick, S., Turkington, T.K., data analysis of weather variables linked to Fusarium head blight
Tekauz, A. et al. (2015) Diversity of fusarium head blight popula- epidemics in the United States. Phytopathology, 109, 96–110.
tions and trichothecene toxin types reveals regional differences in Shah, D.A., De Wolf, E.D., Paul, P.A. & Madden, L.V. (2021) Accuracy
pathogen composition and temporal dynamics. Fungal Genetics and in the prediction of disease epidemics when ensembling sim-
Biology, 82, 22–31. ple but highly correlated models. PLoS Computational Biology, 17,
Landschoot, S., Waegeman, W., Audenaert, K., Van Damme, P., e1008831.
Vandepitte, J., De Baets, B. et al. (2013) A field-specific web tool for Shah, D.A., Molineros, J.E., Paul, P.A., Willyerd, K.T., Madden, L.V. & De
the prediction of fusarium head blight and deoxynivalenol content Wolf, E.D. (2013) Predicting Fusarium head blight epidemics with
in Belgium. Computers and Electronics in Agriculture, 93, 140–148. weather-driven pre- and post-anthesis logistic regression models.
MARD. (2021) Fusarium head blight report-Manitoba agriculture and re- Phytopathology, 103, 906–919.
source development. Province of Manitoba–agriculture. Available Shah, D.A., Paul, P.A., De Wolf, E.D. & Madden, L.V. (2019) Predicting
from: https://​w ww.​gov.​mb.​c a/​agric​ulture/​. [accessed 16th plant disease epidemics from functionally represented weather se-
November 2021]. ries. Philosophical Transactions of the Royal Society B, 374, 20180273.
Matengu, T.T. (2022) Developing risk models to mitigate Fusarium head Spolti, P., Del Ponte, E.M., Dong, Y., Cummings, J.A. & Bergstrom, G.C.
blight in western Canadian cereal production. MSc thesis, University (2014) Triazole sensitivity in a contemporary population of Fusarium
of Manitoba. Available from: http://​hdl.​handle.​net/​1993/​36855​ graminearum from New York wheat and competitiveness of a tebu-
[accessed 7th November 2023]. conazole-resistant isolate. Plant Disease, 98, 607–613.
McMullen, M., Bergstrom, G., De Wolf, E., Dill-Macky, R., Hershman, D., SWDC. (2021) Fusarium resources—Saskatchewan wheat development
Shaner, G. et al. (2012) A unified effort to fight an enemy of wheat commission. Sask wheat development commission. Available
and barley: Fusarium head blight. Plant Disease, 96, 1712–1728. from: https://​saskw​heat.​c a/​fusar​ium-​resou​rces [accessed 16th
Molineros, J. (2007) Understanding the challenge of Fusarium head blight November 2021]
forecasting. PhD thesis, University Park, PA: Pennsylvania State Tamburic-Ilincic, L., Wragg, A. & Schaafsma, A. (2015) Mycotoxin accu-
University. mulation and Fusarium graminearum chemotype diversity in winter
Moschini, R.C. & Fortugno, C. (1996) Predicting wheat head blight inci- wheat grown in southwestern Ontario. Canadian Journal of Plant
dence using models based on meteorological factors in Pergamino, Science, 95, 931–938.
Argentina. European Journal of Plant Pathology, 102, 211–218. van Der Fels-Klerx, H.J., Burgers, S.L.G.E. & Booij, C.J.H. (2010)
Moschini, R.C., Pioli, R., Carmona, M. & Sacchi, O. (2001) Empirical pre- Descriptive modelling to predict deoxynivalenol in winter wheat in
dictions of wheat head blight in the northern Argentinean pampas the Netherlands. Food Additives & Contaminants. Part A, Chemistry,
region. Crop Science, 41, 1541–1545. Analysis, Control, Exposure & Risk Assessment, 27, 636–643.
Musa, T., Hecker, A., Vogelgsang, S. & Forrer, H.R. (2007) Forecasting of Vaughan, M., Backhouse, D. & Ponte, E.M.D. (2016) Climate change
fusarium head blight and deoxynivalenol content in winter wheat impacts on the ecology of Fusarium graminearum species complex
with FusaProg. EPPO Bulletin, 7, 283–289. and susceptibility of wheat to Fusarium head blight: a review. World
Nita, M. (2013) Fungicides–Showcases of integrated plant disease man- Mycotoxin Journal, 9, 685–700.
agement from around the world. https://​doi.​org/​10.​5772/​3251 Wallhead, M. & Zhu, H. (2017) Decision support systems for plant disease
London: InTech. and insect management in commercial nurseries in the Midwest:
Paul, P.A., Lipps, P.E. & Madden, L.V. (2005) Relationship be- a perspective review. Journal of Environmental Horticulture, 35,
tween visual estimates of Fusarium head blight intensity and 84–92.
|

13653059, 0, Downloaded from https://bsppjournals.onlinelibrary.wiley.com/doi/10.1111/ppa.13839 by Cochrane Poland, Wiley Online Library on [26/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
14 MATENGU et al.

Yoshida, M., Nakajima, T., Tomimura, K., Suzuki, F., Arai, M. & Miyasaka,
A. (2012) Effect of the timing of fungicide application on fusarium How to cite this article: Matengu, T.T., Bullock, P.R.,
head blight and mycotoxin contamination in wheat. Plant Disease,
Mkhabela, M.S., Zvomuya, F., Henriquez, M.A., Ojo, E.R.
96, 845–851.
et al. (2023) Weather-based models for forecasting Fusarium
head blight risks in wheat and barley: A review. Plant
S U P P O R T I N G I N FO R M AT I O N
Pathology, 00, 1–14. Available from: https://doi.org/10.1111/
Additional supporting information can be found online in the
ppa.13839
Supporting Information section at the end of this article.

You might also like