Register      Login
Crop and Pasture Science Crop and Pasture Science Society
Plant sciences, sustainable farming systems and food quality
REVIEW (Open Access)

Tactical crop management for improved productivity in winter-dominant rainfall regions: a review

W. K. Anderson https://orcid.org/0000-0003-4940-3102 A E , R. F. Brennan B , K. W. Jayasena C , S. Micic C , J. H. Moore C and T. Nordblom D
+ Author Affiliations
- Author Affiliations

A The UWA Institute of Agriculture, The University of Western Australia, 35 Stirling Highway, Crawley, WA 6009, Australia.

B Formerly: Department of Primary Industries and Regional Development, 444 Albany Highway, Albany, WA 6330, Australia.

C Department of Primary Industries and Regional Development, 444 Albany Highway, Albany, WA 6330, Australia.

D Graham Centre for Agricultural Innovation (alliance of Charles Sturt University and NSW Department of Primary Industries), Albert Pugsley Place, Wagga Wagga, NSW 2650, Australia.

E Corresponding author. Email: wmanderson@omninet.net.au

Crop and Pasture Science 71(7) 621-644 https://doi.org/10.1071/CP19315
Submitted: 26 August 2019  Accepted: 11 June 2020   Published: 16 July 2020

Journal Compilation © CSIRO 2020 Open Access CC BY-NC-ND

Abstract

This study reviews published information on the tactical management decisions needed to maximise economic grain yield in winter-dominant rainfall regions of the Mediterranean type. Tactical decisions are defined as those relating to the period from immediately before sowing to harvest. Tactical management is the principal means by which farmers respond to changing environmental and short-term economic conditions as the season progresses. The review considers published evidence that underpins these decisions and relates to cereal crops (wheat, barley and oats), pulse crops (field pea, faba bean, chickpea and narrow-leaved lupin) and canola.

The criteria used to guide management decisions during the season involve soil and tissue tests for nutrients, knowledge of weed numbers and resistance status in the current and previous seasons, weather conditions that favour disease development, and knowledge of thresholds and biology of insect pests that may warrant control measures. All of these decisions can be related to the timing of the opening rains and the length of the growing season; the crop, pasture or weeds present in the previous two seasons; the presence of pest- and disease-bearing crop residues; and the type of tillage in use. Most of these indicators require further refinement through research across environments, soil types, crop types and production systems.

The likely interactions between tactical or short-term management decisions, longer term or strategic decisions, and genetic factors are discussed. The prevalent use of chemicals in the management of biotic factors that can impact the crops is noted, as is progress towards various systems of ‘integrated’ management of these threats to crop production. Most tactical decisions in rainfed cropping systems appear to be supported by adequate evidence, although some decisions are still based on practical experience and observations.

Application of tactical management practices together with strategic management and use of improved genotypes provides the possibility of achieving rainfall-limited potential grain yield at a regional scale. The papers reviewed have been selected partly on the basis that the experimental treatments achieved the estimated potential grain yield. Where the potential grain yields are not being achieved in commercial crops, it remains unclear whether this is due to inadequate adoption of existing information or inadequate research to identify and address the underlying causes. We highlight the need to devise a simple decision aid to assist farmers and their advisers to respond to the variable seasonal conditions evident since the turn of the Century.

Additional keywords: agronomy, canola, cereals, diseases, economics, insects, legumes, weeds.

Introduction

Grain production in rainfed agriculture is inherently risky, largely owing to substantial variation in amount and distribution of annual rainfall (e.g. Australian Farm Institute 2012), leading to insecurity of the food supply in many regions (e.g. Amede and Tsegaye 2016). However, the grain produced from rainfed agriculture is important for world food supply, and the productivity of the farming systems in these areas needs to be improved to meet future world food requirements (e.g. Sadras and Angus 2006; Anderson 2010; Godfray et al. 2010), especially as the climate changes (Garrett et al. 2006; Ludwig et al. 2009; Luo and Kathuria 2013; Fischer et al. 2014). Grain yields (GYs) in rainfed agriculture can be improved in three broad areas: tactical, short-term or seasonal management (e.g. Turner 2004; Anderson et al. 2005); strategic or longer term management that largely involves soil improvement (e.g. Carter et al. 1998; Bakker et al. 2007); and genetic improvement (Fischer and Edmeades 2010). Analysis of the results of field experiments conducted over a range of rainfed environments tends to suggest that, after accounting for the effects of environment (site, location and season), the impact of management is at least twice that of genotype (Anderson 2010).

Tactical management practices based on field experimentation and observation and applied immediately before and during the growing season may involve adjustments to seeding rates (Leach et al. 1999; Tokatlidis 2014); variations in fertiliser rates (Jarvis and Bolland 1990; Robertson et al. 2012), placement (Nyborg and Hennig 1969) and timing (Anderson 1985; Seymour et al. 2016; Simpson et al. 2016); changes to sowing time and choice of maturity class of cultivars (Anderson et al. 1996; Stephens and Lyons 1998; Sharma et al. 2008) and to seeding depth and method (Schmidt and Belford 1993); dry sowing (Kearns and Umbers 2010); use of herbicide, disease and pest tolerant or resistant cultivars (Jayasena et al. 2018); and various chemical and non-chemical methods of weed management (Farooq et al. 2011; Walsh 2016; Moore and Moore 2020). Tactical management practices and their timing are not a fixed sequence but will vary according to seasonal and market conditions. The evidence discussed here comes largely from the winter-dominant rainfall areas in Australia that are Mediterranean-type environments, classified by Koppen–Geiger as Csa, Csb and BSk (Peel et al. 2007). These environments are found in the Mediterranean Basin and on the west coast of every continent at a latitude ~25–40° with an average annual rainfall of ~250–1000 mm.

Farmers in rainfed agricultural regions have also developed strategies that may need more than one season to take effect, but that can have an impact over many years in improving crop productivity. These strategies may involve practices such as fallow to conserve soil water (Guler and Karaca 1988); integration of animal production with cropping to diversify sources of income (Virgona et al. 2006; Schiere et al. 2006; Bell et al. 2014); use of legume or broadleaf crop rotations in conjunction with cereals to improve soil nitrogen (N) and assist disease and weed management (Donald 1962; Angus et al. 1991; Kingwell 1994; Doole and Weetman 2009; Christiansen et al. 2015; Renton et al. 2015); and various systems involving combinations of practices such as no tillage, retention of crop residues, deep ripping and application of gypsum to address soil compaction (Hamza and Anderson 2003), and controlled traffic farming and crop rotation to maintain soil fertility (e.g. Kingwell et al. 1993; Kingwell and Fuchsbichler 2011; Serraj and Siddique 2012; Loss et al. 2015; Alrijabo 2014; Sommer et al. 2014). There is some evidence that GY responses to tactical inputs and those obtained from strategic treatments such as soil improvements are additive (Anderson 2010). This paper concentrates on tactical management practices that can be used during the short-term, as previously defined.

Although the choice of crop cultivar cannot be changed after sowing, we view the choice of cultivar as a component of tactical management because the decision may change depending on the timing of the opening rains and the likely length of the growing season (Anderson et al. 1996). The use of cultivar mixtures (Fletcher et al. 2019) could be seen as an aspect of tactical management in reducing the risks of such factors as terminal drought and frost damage, anticipated disease risk (Loughman et al. 2000) or weed burden and resistance status. The differential response of crop cultivars to tactical agronomic inputs has not been widely confirmed but may be important under specific circumstances of environment and where differences between genotypes are large (Anderson et al. 2011).

Despite the difficulties of estimating and verifying the potential GY for rainfed crops (Anderson 2010), there is some evidence that the best farmers in the more favoured areas (fertile soils, more evenly distributed and reliable rainfall) and/or using the most advanced management can approach the estimated potential for common wheat (Triticum aestivum L.) (Poole et al. 2002; Abeledo et al. 2008; Richards et al. 2014; Robertson et al. 2012; Anderson et al. 2016). However, when regional average GYs are considered, the tendency is for GYs to approach the estimated potential only at seasonal rainfall ≤~250 mm (Anderson et al. 2005). This probably suggests that the management techniques associated with low seasonal rainfall are well accepted and dealt with by farmers, whereas the risks associated with the opportunities to increase GYs in the wetter seasons are unacceptably high, given the difficulty of predicting seasonal conditions. In either case, an estimate of the gap between average or actual GY and the estimated potential can be useful in deciding whether further effort to improve GYs is possible and likely to be profitable (e.g. Keating et al. 2003; Hochman et al. 2012).

In any given growing season, there is a sequence of decisions that are influenced by the likely GY or target GY, which is determined largely by the amount and timing of the seasonal rainfall. The timing of the opening rains sets the likely length of the growing season, the likely amount of rainfall, and thus the GY that is possible for a given set of soil physical and chemical conditions (Anderson 2010). Decisions regarding weed, pest and disease management during the season will be based on the likelihood of these factors reducing the target GY. The extent to which the target GY is realised will depend largely on the decisions made to manage and defend crop growth as the season develops. The whole-farm implications of various decisions have been examined further with respect to output responses (Kingwell 1994; Kingwell and Pannell 2005) and risk aversion (Kingwell 1997).

This paper summarises the published information available for making tactical management decisions and demonstrates how this information can be used to allow crop managers to approach more closely the target GY as set by the seasonal rainfall. Most of the published information considered in this review contains data that compare experimental GY with some estimate of the potential GY for each growing season. The information on tactical management, along with information on strategic management and improved genotypes, underpins the advice commonly given to farmers and, in addition, can supply functions in the construction of crop models.


Tactical agronomy

Start and length of the growing season

Most farmers have developed an experiential rule about the amount and timing of the opening rains constituting the start of the growing season. This rule commonly describes an amount in a single fall after a pre-determined date that may be based on the likelihood of frost damage resulting from early sowing. For example, 20 mm rain over 1–2 days after April 25 is a common criterion in some parts of the Western Australian (WA) grain belt. These ‘rules of thumb’ can likely be strengthened by further research. The end of the effective growing season can be estimated as the average temperature associated with grain maturity. For example, 23°C average maximum for wheat as used by French and Schultz (1984).

With these criteria, the length of each growing season for a given location, using local records, can be estimated and this length in turn can be related to the likely amount of seasonal rainfall each year (Anderson 2010). Thus, in a given season on the date when the opening rains are received, the likely length of the season and the likely amount of rain can be estimated for a given location.

Target grain yield

By using the estimated seasonal rainfall, a target GY can be estimated at the start of each season with either a simple water-balance equation (Nix and Fitzpatrick 1969; French and Schultz 1984; Anderson 1985) or a more sophisticated crop model using more variables that may influence crop growth and GY (e.g. Stephens et al. 1989; Keating et al. 2003). This target GY, however estimated, can be updated during the season according to the actual rainfall received, as distinct from the estimate at the beginning of the season. This may be one method of adapting to the changes in rainfall patterns that have been noted by Stephens et al. (1989) and reflected in average wheat GYs subsequently (ABARE 2017). It may also be used to vary the rates of fertilisers required to achieve the target GY.

Time of sowing and choice of cultivar

It is generally accepted that longer season (later maturing) wheat cultivars should be sown early in the season when the opening rains occur early, and shorter season (early maturing) cultivars can be sown later (e.g. Doyle and Marcellos 1974; Anderson et al. 1996; Stephens and Lyons 1998). Long-season wheat cultivars may yield the same as mid- or short-season types when sown later (Sharma et al. 2008); however, the grain quality (small grain screenings, hectolitre weight) of long-season cultivars may be unsatisfactory when sown later in the season (Sharma and Anderson 2004).

In addition to reducing exposure to drought and frost damage by matching cultivar maturity to anticipated length of growing season, seasonal financial risk can be reduced by choosing cultivars attracting a price premium. For example, hard, soft and high starch-quality wheat cultivars that may qualify for premiums can be grown with minimal risk if the rotation and soil type are appropriate, even if their GY potential is not high (Anderson and Sawkins 1997; Anderson et al. 1995, 1997).

Plant population or seed rate

The target GY can be used to estimate the plant population or seed rate needed to support that target. For example, a population of 40 wheat plants/m2 is required in WA for each tonne of target GY (Anderson et al. 2004). However, given the benefits of increased plant numbers for weed management (Radford et al. 1980; Walsh and Minkey 2006) and the reduced crop establishment expected at higher seed rates (Del Cima et al. 2004), the seeding rates required for various GYs of wheat, for example, can be estimated as in Table 1. It can be hypothesised that if a seed rate of <100 kg/ha is used for a target GY of wheat of 4 t/ha, then plant numbers could be the factor that limits GY in a given season. In addition, information from the current and previous seasons about the likely weed burden can be used to adjust the seeding rate.


Table 1.  Estimated plant populations and seed rates for wheat according to target grain yield
Row spacing of 20 cm and average seed size of 35 mg are assumed for the calculations
T1

This type of relationship also appears to apply to narrow-leaved lupin (Lupinus angustifolius L.) in WA (French et al. 1994) and canola (Brassica napus L.) (Seymour 2011; French et al. 2016; Roques and Berry 2016). This requires further testing for a range of crop species relevant to the rainfed agricultural regions because similar relationships may not hold in other crops (Jettner et al. 1998). In all cases, the germination percentage and average size of the seed to be sown should be factored into the seed-rate decision (e.g. French et al. 1994).


Determination of fertiliser type and amount

For nutrients such as phosphorus (P), potassium (K) and sulfur (S), soil tests taken before seeding are reliable indicators of deficiency for plant growth and thus of the need for application in the current season (Peverill et al. 1999; Fageria 2009). The micronutrient status of the current crop is best assessed by using plant tissue tests (Reuter and Robinson 1997), with micronutrients applied as recommended (e.g. Graham 2008). If plant-tissue test results can be returned quickly enough, these micronutrients can be applied as foliar sprays during the season, thereby avoiding GY loss. Tests for P, S and N can also be made and corrective fertilisers applied, particularly for K and S (Anderson et al. 2015). However, various commercial models are available for in-crop advice on strategic applications of fertiliser. Commercial fertiliser companies and government advisers use various (in-house) models mainly for N applications (e.g. SyN and NPDecide) in WA, and various models based on APSIM (Keating et al. 2003) are available for private agronomists and cropping system advisors. The use and effectiveness of crop models needs to be more widely and objectively assessed.

Tactical foliar application of nutrients

It is generally better to ‘feed the roots’ rather than the foliage, although a few exceptions exist depending on soil conditions (e.g. Ali et al. 2016). In order to clarify the role of foliar-applied nutrients as a means of tactical management, it is important to look at the crop’s entire nutritional requirements (Franke 1967).

Foliar applications should not be relied on to correct macronutrient deficiencies in rainfed cereal crops, because soil-applied nutrients frequently give higher GYs (Seymour and Brennan 1995; Abdul et al. 2012; Fernández and Brown 2013).

Phosphorus

Foliar application of P is not recommended anywhere in the world, partly because the quantity of P needed, 10–20 kg/ha (see Table 2), makes it difficult to apply as a spray. Chemicals providing <5 kg P/ha rarely correct P deficiency. In addition, some studies have shown that, although the timing of foliar application with inclusion of a surfactant in the P solution is important for absorption of P by wheat leaves, the improved foliar uptake does not lead to an increase in GY (Fernández et al. 2014; Peirce et al. 2019). Because P is required early in growth and broadcast applications are frequently about half as effective as drilled applications, fertiliser is best drilled at sowing (Jarvis and Bolland 1990). The decision is thus a tactical one made at or immediately before sowing.


Table 2.  Range of concentrations (kg/t) of nutrients found in wheat grain and straw in Western Australia
Click to zoom

Nitrogen

Foliar application of N to cereals has attracted considerable attention (Reeves 1954; Finney et al. 1957; Gooding and Davies 1992), particularly applications close to flowering, which increase the number or size of grains in the ear (Bly and Woodard 2003). Urea (CH4N2O), urea–ammonium nitrate (NH4NO3) solution mixture (UAN), ammonium sulfate (NH4SO4), and ammonium nitrate can be applied as spray solution (Fageria et al. 2009). However, the amount applied for tactical N applications is limited to the maximum N concentration that can be used without freezing pipes and nozzles, rather than the chemical solubility of the N compounds (~330 g urea/L water, 500 g ammonium nitrate/L water). UAN is a solution containing 420 g N/L water. There is little difference in GY when different N sources are compared on the same rate of N applied (Fageria et al. 2009). Generally, under rainfed conditions with terminal drought, low soil moisture during grain-filling limits the potential for increased GY from foliar N application; however, increases in protein content of wheat grain after late application of foliar N have been reported (Reeves 1954). In general, the tactical (or split) application of granular fertiliser can be an effective means of tactical management of N on neutral to acid soil types (see below).

Potassium

If K deficiency is diagnosed by either plant tissue analysis or visual symptoms, K fertiliser can be applied to the soil surface 4–6 weeks after sowing. Foliar sprays generally do not supply sufficient K to overcome a severe deficiency and can scorch leaves of crops in some circumstances. Foliar-applied K, along with other micro- and macronutrients, did not increase the GY of narrow-leafed lupin in WA (Seymour and Brennan 1995).

Some field experiments have evaluated the efficiency of K applied as a foliar spray versus a soil application (John and Lester 2011; Abdul et al. 2012; Ali et al. 2016; Farooqi et al. 2012). A foliar spray of 3% K was more efficient for increasing growth and GY than soil application and fertilisation (Ali et al. 2016).

Sulfur

When S deficiency is observed or diagnosed by plant analysis, the usual practice is to apply S fertiliser to the soil surface (Brennan and Bolland 2008). In a 2-year study, Ozanne and Petch (1978) measured GY increases of 22% following foliar fertilisation with a mix of N, P, K and S. However, Seymour and Brennan (1995) found that 4 kg S/ha applied 2 weeks after flowering reduced the GY of narrow-leaved lupin by ~10% (300 kg/ha), despite no evidence of foliar damage. Limited data are available on the effect of S foliar spray for the other crop species grown in rainfed agricultural regions.

Copper (Cu)

Foliar application of Cu can be useful up to flowering if a deficiency has been diagnosed in growing plants (Gartrell 1981). Foliar spray usually contains 10–15 g CuSO4.5H2O/L water. A single soil application of Cu at recommended rates can provide a longer residual effect for the growing crop, whereas foliar applications generally require consecutive applications. However, foliar Cu sprays are seen as an appropriate response if plant-tissue analysis during the growing season has indicated a Cu deficiency (Gartrell 1981; Brennan 2000). Where crops are grown on subsoil moisture, the topsoil where the Cu fertiliser is placed is frequently dry and the fertiliser is unavailable to root uptake, in which case foliar applications are more effective (Gartrell 1981; Grundon 1980).

Zinc (Zn)

Severe Zn deficiency can affect young seedlings, usually in the first 3–4 weeks following emergence (Brennan 1991). Foliar application of Zn helps the remaining seedlings to survive, even if a large percentage of the spray lands on the soil surface. Effects on GY are minimised if foliar-applied Zn is applied immediately symptoms are seen. In situations when Zn deficiency is less severe, foliar sprays are required as soon as symptoms appear or after plant-tissue analysis indicates Zn-deficiency problems.

Manganese (Mn)

In some situations, a foliar spray of Mn is the most effective treatment, particularly on highly alkaline soils (pH >8.5) where Mn is less soluble (Gettier et al. 1985). In acid soils, for example in WA, sporadic Mn deficiency occurs in gravelly soils, particularly in drier growing seasons (Smith and Toms 1958; Brennan and Bolland 2011). Patches of crop displaying Mn deficiency can be sprayed with MnSO4.5H2O, which is a relatively cheap and effective source of Mn.

Manganese is required for the developing seed in pods of narrow-leaved lupin and other grain legumes as the crops mature. During this part of the growing season, the topsoil is frequently dry, and soil-applied Mn is unavailable for uptake by roots. Hence, a tactical Mn foliar spray is effective to overcome Mn deficiency in lupin if the plant is sprayed when the seed pods on the main stem are ~2.5–3.0 cm long. If delayed beyond this time, Mn spraying is less effective (Seymour and Brennan 1995). Manganese deficiency results in split and shrivelled seed within the pods, resulting in GY losses and decreased profits for producers.

Molybdenum (Mo)

Usually fertiliser application of 75–100 g Mo/ha lasts ~10 years in acidic sandplain soils (Doyle et al. 1965; Brennan 2006). If Mo deficiency is diagnosed by plant-tissue analyses in crops such as wheat, barley (Hordeum vulgare L.) or canola, a solution of sodium molybdate (Na2MoO4, 39% Mo) or ammonium molybdate ((NH4)6Mo7O24, 54% Mo) applied to the leaves of the plants corrects the deficiency when applied at rates of 0.20–0.40 g ammonium or sodium molybdate/L water. For example, tactical application of foliar Mo sprays at mid-flowering of 40 g Mo/ha overcame Mo deficiency for canola grain production (Brennan and Bolland 2011). However, higher foliar rates of 50–150 g/ha of the Mo sources are recommended for some crop species (Bell and Dell 2008).

Liming (CaCO3) may be required to ameliorate soil acidification in some regions, and this will also alleviate Mo deficiency induced by soil acidification if sufficient lime is added to raise the pHCa of the top 10 cm of soil to ≥5.5 (Brennan and Bolland 2011). If deficiency of Mo is diagnosed for crops grown on acidic soils, a foliar spray is fully effective if applied at ~20 g Mo/ha.

Boron (B)

Boron deficiency is rare in wheat and other cereals (Brennan et al. 2015). In addition, species with a higher requirement for B (e.g. canola and lupin) are usually adequately supplied in the soils of WA, even where soil extractable B levels are low (~0.003 g B/kg in CaCl2 extractant). However, the redistribution of foliar-applied B to actively growing young tissues that require B for growth is limited in most crop species, except lupin. Thus, foliar application of B is of limited use for most crop species even if B deficiency is diagnosed. This suggests that critical soil-test concentrations for crop production need to be defined.

If B deficiency is diagnosed by plant-tissue analysis, B can be sprayed at 0.1–0.5 kg B/ha at the flag leaf stage (Z37–41; Zadoks et al. 1974) to increase GY by reducing pollen sterility (Dell et al. 2002). However, application of B fertiliser, particularly borax (Na2(B4O5(OH)4).8H2O), can reduce germination and seedling density (Brennan et al. 2015).

Removal of nutrients

One guide for determining rates of fertiliser application needed in the current season is to estimate the likely amount of each nutrient removed in the grain (for wheat, see Table 2; after Gartrell and Bolland 2000). The target GY can then be used as the starting point for estimating fertiliser application rates. Soil type, previous crops and seasonal conditions need to be considered, and local experience at the farm or paddock level, in combination with soil and/or tissue tests, is often used to ‘fine-tune’ these decisions (e.g. the use of test strips when applying nutrients diagnosed as potentially deficient). Adoption of variable-rate fertiliser technology (Robertson et al. 2012) can be a further aid in optimising costs and returns where several sources of information need to be considered.

Tactical N application

Several forms of N have been assessed for tactical application in field experiments in WA, including urea, ammonium nitrate, ammonium sulfate, calcium ammonium nitrate (variable formulations) and anhydrous ammonia (NH3) (Mason 1968, 1977). The composition of these N fertilisers and current costs should be compared on cost per unit N (Mason 1968, 1977); however, urea is usually the most cost-effective N source for tactical applications. The main benefit of N is to increase the number of ears per ha in cereals (Simpson et al. 2016), and the greatest benefit is achieved by applying the N early, especially in short-season environments, so that tillering can be increased (Littler 1963; Ellen and Spiertz 1980). However, the optimum tactic recommended for timing of N applications to annual crops varies, especially according to soil and rainfall conditions (Nordblom et al. 1985; Ladha et al. 2005). Application of N at sowing appears preferable where the crop is grown largely or partially on stored soil water (Cooper 1974). Application before expected rainfall in addition to a dose of N at sowing also appears advantageous where the soils are unlikely to be prone to nitrate leaching (Anderson 1985; Angus and Good 2004; Simpson et al. 2016). In temperate environments with cold winters and relatively reliable precipitation, an appropriate tactical management is to apply the N fertiliser according to the growth stage of the crop (Ellen and Spiertz 1980; Christensen and Mients 1982; Tosti et al. 2016). The N status of the crop can be compared with ‘adequate’ levels from tissue tests as described for local conditions (e.g. Reuter and Robinson 1997; Simpson et al. 2016). In the high-rainfall zone of the south of WA, where the topsoils are often sandy and the rain is concentrated in the growing season, it may be appropriate to replace losses due to leaching- and/or waterlogging-induced denitrification after heavy rain events (Simpson et al. 2016). Estimates of crop N status in the field have been developed by using methods involving crop-canopy reflectance, leaf transmittance and chlorophyll concentration (Muñoz-Huerta et al. 2013). Many commercial, ground-based methods (i.e. Yara N-Sensor, GreenSeeker, CropScan) and satellite-mounted sensors (i.e. QuickBird) measure crop-canopy reflectance in the visible and/or infrared wavebands for the purpose of estimating N in plants (Li et al. 2010; Vigneau et al. 2011). There is a paucity of data for rainfed crops, and this study suggests that further research is required. The application of tissue tests as an aid in assessing fertiliser requirements can be complicated by competition from weed species such as annual ryegrass (Lolium rigidum Gaud.) (Palta and Peltzer 2001).

From a tactical management perspective, the total amount of N required to satisfy the target GY can be estimated at sowing. For example, using information from Gartrell and Bolland (2000), a 4-t crop of wheat would contain 64–104 kg N in the grain and a further 8–40 kg in the straw (Table 2). Part of the total N required can then be applied at sowing, and the remainder applied according to the preferred tactics depending on the soil and weather conditions. One advantage of this application system is that the later applications can be withheld if seasonal conditions do not meet expectations and the probability of reaching the target GY is reduced (Seymour et al. 2016; Simpson et al. 2016), or if the chance of ‘haying off’ is increased (van Heerwaarden et al. 1998).

Use of tillage in tactical management

Full (mechanical) tillage, which is a tactical means to address short-term issues, is considered one of the oldest methods for controlling weeds, pests and diseases (Yenish et al. 1992; Horne and Page 2008; Franzluebbers et al. 2011). Disturbance of the soil profile and the removal (by incorporation) of surface plant residues eliminates suitable shelter and host plants for pests (e.g. snails) and reduces the risk of stubble-borne diseases such as net-type net blotch (Pyrenophora teres f. teres) and spot-type net blotch (P. teres f. maculata) (Jayasena and Loughman 2001; Horne and Page 2008; Le Gall and Tooker 2017). Tillage distributes weed seeds throughout the tilled soil layer, with the highest concentration of weed seeds generally in the 0–5 cm soil layer (Franzluebbers et al. 2011). Seeds of common weeds buried deeper than 10 cm often germinate if sufficient soil moisture is present but they usually fail to emerge. Tillage can stimulate mass germination of weeds such as annual ryegrass and wild oats (Avena fatua L.), allowing control of these seedlings with a single herbicide application before crops are sown (Madin 1993; Franzluebbers et al. 2011). Full tillage is no longer practiced in many rainfed agricultural areas (e.g. Llewellyn and D’Emden 2009; Yigezu et al. 2014; Loss et al. 2015), although it does have a place in treating soil compaction (Hamza and Anderson 2005) and managing certain weed species (see below), and it may be used to achieve the above benefits.

Currently, farming systems in the south of WA use minimum or no tillage techniques, and ≥30% of plant residues are retained on the soil surface to reduce water and wind erosion (Derpsch 2003). However, retention of crop residues in rainfed cropping systems in Australia and elsewhere does not appear to have had unequivocal benefit for crop GYs (Scott et al. 2010; Loss et al. 2015). We conclude that further research is required to delineate clearly the separate effects of reduced tillage and residue retention and to establish the appropriate conditions for full soil disturbance across a range of cropping situations.

The adoption of minimum or no tillage and stubble retention as part of a system of conservation cropping can increase the disease load in paddocks, where some diseases such as such as spot-type net blotch of barley can persist on stubbles such that >2 years is required to reduce the disease burden of the stubble effectively (Jayasena and Loughman 2001). If seeding occurs in ungrazed pastures or if the preceding cereal GY was >1 t/ha, then tillage may still be an option. According to Leonard (1993), cereal paddocks with these GYs are likely to have residues >1.4 t/ha, which contributes to poor crop establishment in the following year. On the positive side, tilling the paddock requires no removal or management of stubble and may lead to better crop establishment. The decision to use deep tillage (inversion ploughing or deep ripping) has often been made by farmers in WA after testing for soil compaction by using an improvised, pointed steel rod.

The impact of tillage systems on soil compaction and its treatment across common soil types has been studied in WA (Hamza and Anderson 2003), with the roles of deep ripping and gypsum application measured. There is probably a further requirement to examine the effects of these strategic practices and their possible interaction with the tactical practices discussed here.

Tactical applications of crop rotation

Crop rotations in which a pasture, pulse or oilseed crop is grown after a cereal crop can significantly influence a range of factors including changes in soil N and management of pests, weeds and diseases. Fixed rotational systems are rarely used, with the final choice influenced by several considerations including short-term fluctuations in commodity prices for grains and livestock (Lawes 2015; Lawes and Renton 2015). If consideration is given to the value of grain from the different crops that are compared in this paper, canola had the highest average price in 2019, at AU$601/t, followed by lupins ($430/t), wheat ($309/t) and barley ($233/t) (Grain Central 2019; Cargill 2020a, 2020b, 2020c). Given the relative prices and expected grain yields of the crops, it is often profitable to have canola in the cropping rotation despite its susceptibility to pests, increased input costs and environmental risks (Kirkegaard et al. 2016).

Decisions regarding crop sequences can be made immediately before seeding, but some of the inevitable interactions with the longer term strategic decisions are worth discussing. Canola is often grown as the first crop after a pasture to control weeds and diseases before the cereal phase (Kirkegaard et al. 2016). However, pastures sustain higher numbers of pests such as earth mites and weevils that can damage the following seedling canola crop, leading to increased insecticide usage (Micic 2005; Horne and Page 2008). In recent years, the frequency of canola has increased in the rotation, with a similar decrease in pastures, especially in WA. This increase in canola has reflected its profitability as a commodity in its own right as well as a strategy to control weeds and diseases for the following cereal crops (ABARE 2017). The reduced time between canola crops has led to an increase in stubble-borne diseases such as blackleg (Leptosphaeria maculans), increases in applications of N and of herbicides to control weeds in the canola phase (see Kirkegaard et al. 2016), and increased insecticide usage (Macfadyn and Hill 2017). Where the goal is to maximise profits, there is a need to consider an integrated approach for the management of disease, weeds, pests and fertiliser applications (Brennan 1989; Bockus and Claassen 1992; Doole and Weetman 2009; Nichols et al. 2009, 2012).

In WA, disease risk for barley is increased if the crop is grown at intervals of <4 years, but it is more profitable to grow barley with a 2–3-year rotation even if stubble-borne diseases such as spot-type net blotch are an issue (Jayasena and Loughman 2001). Currently, the disease risk for short rotations is high, and this is managed through the use of fungicides to decrease stubble-borne diseases. However, repeated applications of fungicides increase the risk of resistance developing in the disease pathogen. For example, malting varieties of barley have poor resistance to powdery mildew (Blumeria graminis f. sp. hordei), but owing to market demand, these varieties are most commonly grown with triazole-based seed-dressing fungicides. The pathogen has now developed resistance (Tucker et al. 2015), and this can reduce GY by up to 40% if infection occurs before flag-leaf emergence and up to 25% if infection occurs after this time.

Management of the ‘green bridge’

Mixed-farming enterprises commonly leave volunteer plants and weeds intact for grazing by livestock during the pre-crop period (hence the ‘green bridge’). This practice has implications for insect, disease and weed management (Coutts et al. 2018). Feeding of livestock on green weeds and crop stubble can decrease (but not eliminate) disease inocula of, for example, air-borne diseases such as barley leaf rust (Puccinia hordei). However, this practice can lead to a decrease in the amount of standing stubble, leading to germinating crops being more susceptible to colonisation by aphids (Jones 1994; Coutts et al. 2015).

The success of herbicide applications in controlling a green bridge depends on the weed seedbank that may germinate with subsequent rainfall events, as well as the size of the weeds present at emergence (Walsh et al. 2004; D’Emden and Llewellyn 2006; Bastiaans et al. 2008). In addition, a green bridge depletes the available stored soil moisture for the subsequent crop (Walsh et al. 2004; Bastiaans et al. 2008). Application of non-residual herbicides may be needed after each germination of weeds; residual herbicides are not commonly used for summer weed control.

The timing of herbicide applications to control the green bridge is important for reducing numbers of pests and can be as effective as an insecticide application (Macfadyn and Hill 2017). However, the optimum time to apply herbicides is ≥14 days before seeding (Coutts et al. 2015). This will lead to a fallow period of at least 10 days before crop emergence, thereby reducing pest numbers. Fast-acting herbicides may be required where timings are tight so that the fallow period is sufficient to stop pests such as aphids or mites transferring from dying plant hosts onto germinating crops.


Weed management

Weed-management decisions are driven mainly by the economic impact of the weed on the current crop, the impact it may have on the following crop, and the propensity for the particular species to develop resistance to common herbicides. The impact on the current crop can be due to competition (e.g. annual ryegrass, Moore 1979), reduction of available soil moisture (e.g. summer weeds), or grain contamination (e.g. by seeds of wild radish (Raphanus raphanistrum L.) or sclerotium of ergot of rye (Claviceps purpurea)). Thresholds are used to determine when control is required. Weed control based on the impact on following crops is driven by the fact that it is often easier and less costly to control grasses in broadleaf crops, and broad-leaved species in cereal or grass crops. Thus, higher levels of control of wild radish in cereals may be practised if a legume or Brassica crop is following. Similarly, high levels of grass control may be practised in broadleaf crops if a cereal crop is planned for the following year. Resistance is another important factor in decision making, especially for annual ryegrass, brome grass (Bromus rigidus Roth.) and wild radish. In these cases, growers often opt for maintaining very low levels of the weed in continuous cropping fields in order to reduce the risk of developing resistance to the most economic herbicides.

Overall, the weed species in agricultural ecosystems in rainfed temperate environments tend to be a mix of species of Poaceae (grasses), Asteraceae (daisies and thistles), Brassicaceae (radish, turnips and mustards) and Fabaceae (clovers, medics and grain legumes) (Moore and Wheeler 2020). Profitable crops also tend to come from these families, with wheat, oats (Avena sativa L.) and barley from the Poaceae family, canola from the Brassicaceae, and lupins, peas and beans from the Fabaceae. Because these crops are often rotated, the major weeds are often volunteers from the previous crop in addition to the naturalised species (green bridge). In Australia, native plant species are rarely weeds of winter crops but can occur as summer weeds (Moore and Moore 2020).

Because elimination of all weeds and their seeds seldom occurs in the previous growing season, tactical weed management is essential for each growing season and for each crop species. This includes cultural treatments such as cultivation, burning, weed-seed collection at harvest, grazing, rotation and fallow, in addition to herbicide treatments (Dodd et al. 1993; Seymour et al. 2012; Walsh et al. 2017). Insofar as weed management has a tactical component, it can be guided by knowledge of the weed population in the year before cropping and by observation of weed-seedling emergence after the opening rains in the current season (Dodd et al. 1993).

Growers may wait for emergence of weeds before spraying with non-selective (knock-down) herbicides if there is an early break to the season. This application may be repeated in 3–14 days with a follow-up spray before sowing the crop. If there is a late break to the season, growers may sow the crop before the rains arrive (dry seed) and apply residual herbicides before sowing or after sowing and before crop emergence. Selective herbicides can then be used once the crop has established. In these situations, growers may also opt to switch to herbicide-tolerant crop varieties so that glyphosate, imidazolinones or triazines may be applied to the crop after emergence. The particular herbicides used depend on the crop, weeds, tillage system (as some herbicides require incorporation), and resistance status of the weeds—in particular annual ryegrass, which has widespread resistance to the Group A (‘fop’ and ‘dim’ grass-selective) and Group B (sulfonylurea broad-spectrum) herbicides. Typical mixtures and sequences of herbicide application are summarised for the major crops in ‘Favourite Brews 2020.pdf’ in Moore and Moore (2020). In most cases, this will include herbicides that provide good control of weeds from the Asteraceae, Brassicaceae, Fabaceae and Poaceae families. Computer models such as HerbiGuide and HerbiRate are used tactically to adjust herbicide selections or rates of application on the basis of local conditions (Moore and Minkey 1997; Moore and Moore 2020). Several other models are rarely used by growers but are good educational aids to demonstrate principles of weed control and agronomic interactions (e.g. Legizamon et al. 1980; Grundy and Mead 1998; Monjardino et al. 2003; Jones et al. 2005; Somerville et al. 2018).

Biological control agents have been released against several weeds with varying effects. Common heliotrope (Heliotropium europaeum L.), doublegee (Emex australis Steinh.), fiddle dock (Rumex pulcher L.), Paterson’s curse (Echium plantagineum L.), skeleton weed (Chondrilla juncea L.) and various thistles are agricultural weeds that have been targets for bio-control, using a range of agents including rusts, beetles, mites, midges, moths, flies and weevils (Dodd et al. 1993). Because their level of bio-control varies from season to season, tactically applied supplemental control is often used. This area needs further investigation.

Minimum tillage is the most common planting practice in broadacre, rainfed cropping systems in Australia (Anderson et al. 2005). This can be related to various combinations of increasing costs of cultivation; decreasing costs of herbicides, insecticides and fungicides; larger farm sizes and machinery; risks of erosion; and difficulty in attracting labour during peak periods. However, it may complicate weed management.

Management practices

Tactical practices may include:

  • Changes in weed control in leguminous crops and pastures based on price movements of N or grains.

  • Changes in weed control so as to vary the carryover of weed seed into the following season, based on changes in predicted relative prices of crops versus livestock in the following season where crop weeds are the main pasture species.

  • Timing of weed control, with early control generally giving higher GY than late control. Seasonal conditions will influence timing and products used for early control, and late germinations of weeds that cause grain contamination may necessitate a later spray.

  • Herbicide rate adjustment to take account of current environmental conditions (Minkey and Moore 1996) or weed density (Moore and Moore 2020).

  • Use of a ‘base recipe’ for the crop (Moore and Moore 2020) or advice from a consultant. Many growers then adjust this for tactical management of specific weeds or where herbicide resistance occurs. The base recipe usually contains herbicides or practices that control the major families of weeds to reduce the risk of minor weeds predominating in response to the weed control implemented.

Both pre-emergent and post-emergent herbicides are extensively used as are residual and non-residual herbicides in the Australian rainfed cropping systems (Moore and Moore 2020).

Pre-seeding and pre-emergent practices

In addition to practices of summer weed control and reduction of the green bridge to conserve moisture and reduce effects of disease, insect and allelopathy, several tactics are imposed close to seeding.

Growers generally wait for a germination of weeds before spraying with glyphosate, or less commonly, they use paraquat or paraquat + diquat mixtures as the knock-down herbicide. Trifluralin is often added as a pre-emergent herbicide before seeding common crops such as cereals, canola or legumes. Atrazine and simazine are commonly used residual herbicides before seeding lupins and triazine-tolerant canola. Imidazolinone herbicides are used on Clearfield varieties of cereals and canola, and various speciality herbicides are used in crops such as chickpea (Cicer arietinum L.), faba bean (Vicia faba L.) and lentil (Lens culinaris Medik.) (Moore and Moore 2020). In paddocks with a high weed burden, a shallow cultivation may be used to encourage the weeds to germinate, thereby allowing greater control with the knock-down herbicides.

If the rains at the start of the season are ‘late’, the tactics used include seeding crops into dry soils with knock-down herbicides, replaced by residual herbicides such as atrazine for triazine-tolerant varieties of canola or imidazolinone herbicides for Clearfield or imidazolinone-tolerant varieties. Cereals and conventional varieties of canola may be sown following the application of trifluralin herbicide, and a tactical decision may be made to add a more soluble product such as metolachlor or to top-up after the rains. Where herbicide resistance is found, active ingredients such as pyroxasulfone (Sakura, Bayer CropScience) may be substituted for trifluralin before seeding cereals (Moore and Moore 2020).

Post-emergent herbicides

Various post-emergence selective herbicides are commonly used in cereals, and many contain MCPA (a Group I herbicide) mixed with products such as bromoxynil, diflufenican, picolinafen, terbutryn, pyrasulfotole, linuron or diuron, from other herbicide groups (Moore and Moore 2020). The choice of herbicide is usually a tactical decision based on the species and densities of the weeds present at the two-leaf to tillering stages of the crop.

Hormone herbicides such as 2,4-D are still widely used, whereas the use of Group B herbicides has decreased as herbicide resistance in the major weeds (i.e. annual ryegrass and wild radish) has developed (e.g. Owen et al. 2014). Most growers will inspect the crop at 6–8 weeks after germination and make a tactical decision on the need for follow-up herbicide application based on the degree of weed control achieved and the level of late germinations.

Many growers have management practices in place to reduce the risk of herbicide resistance developing. These practices include rotating between herbicide groups, using non-chemical methods of weed control, maintaining low weed densities, sowing clean seed, and actively controlling patches of weeds that appear to be resistant. About three-quarters of annual ryegrass populations are resistant to commonly used Groups A and B herbicides (Owen and Powles 2018). In paddocks with high weed burdens, glyphosate-tolerant canola varieties may be planted and glyphosate applied post-emergence. Failures of weed control due to resistance are usually remedied with tactical application of an alternative herbicide, or a decision to cut the crop for hay rather than allowing it to progress to grain harvest.

Non-chemical weed control

Non-chemical methods for weed management include:

  • Tactical shallow cultivation to encourage weed germination when the break of the season occurs at least 1 week before the planned planting date.

  • Inversion (mouldboard) ploughing in areas with high levels of herbicide resistance, or in combination with other practices such as the redistribution of nutrients and lime or reducing non-wetting.

  • Concentration of chaff and harvested weed seed into narrow windrows behind the harvester.

  • Tactical burning of stubbles or harvest windrows when conditions allow.

  • Weed-seed collection at harvest by using chaff carts or stubble balers, based on the weed-seed set in the paddock or price of stubble.

  • Harvest weed-seed control, using various machines that damage seed as it leaves the harvester (Walsh et al. 2017). These are turned off in areas that are relatively weed-free.

Herbicide resistance

Owen and Powles (2018) have reported high levels of resistance to Groups A and B herbicides in ryegrass and wild radish in many cropping areas of Australia. Populations resistant to other commonly used herbicide groups also occur, but at very low levels. Tactical decisions are taken to manage these populations with other herbicides or control methods (Dodd et al. 1993). Despite high levels of resistance, growers are maintaining adequate levels of weed control (Owen and Powles 2018). Many growers have adopted practices that are applied tactically over the cropped areas to minimise the risk of herbicide resistance.

Other factors influencing weed control decisions

Tactical control of particular weeds is practiced for a variety of other reasons in rainfed agricultural systems including:

  • Control of capeweed (Arctotheca calendula (L.) Levyns), thereby reducing the effects of redlegged earth mite (Halotydeus destructor (Tucker)) and vegetable weevil (Listroderes difficilis (Germain)) on canola.

  • Control of small crumbweed (Dysphania pumilio (R.Br.) Mosyakin & Clemants) and stinkwort (Dittrichia graveolens (L.) Greuter) to reduce allelopathy in following crops (Moore and Moore 2020).

  • Summer weed control to conserve soil moisture (Bastiaans et al. 2008), reduce disease levels, and reduce insects including caterpillar pests such as pasture webworm (Hednota spp.), mites and molluscs.

In conclusion, we argue that growers are managing weed populations to levels that are not having a major impact on GY. Further research on optimising weed control would lead to greater profitability, because many weed-control decisions are driven by risk aversion rather than cost effectiveness. Biosecurity is an area that also requires ongoing support to protect the industry from new threats such as Star of Bethlehem (Ornithogalum umbellatum L.), three-horned bedstraw (Galium tricornutum Dandy), skeleton weed (Chondrilla juncea L.) and other undetected species. Although there are policies to control new invasive species, rapid tactical responses are required on discovery. Market access also needs continuing research to ensure that residue levels of herbicides and pesticides are kept at acceptable levels and other contaminants such as ergots and ryegrass toxicity are minimised.


Insect pests

Tactical management of insect pests relies on frequent field observation in the current season (counts of invertebrate numbers and/or damage assessments to plants) when weather conditions are likely to favour the pests, and use of control measures when thresholds are reached. Some thresholds can be related to environmental conditions that favour multiplication of some insect species (Nichols et al. 2009).

Current pest-management practices in commercial systems

Tactical control of agricultural pests in Australian pastures and cropping environments relies heavily on pesticide applications (Micic et al. 2008; Umina et al. 2011), which are routinely applied as a prophylactic spray for the control of pests regardless of pest densities (James 2000; Macfadyn and Hill 2017). In WA, prophylactic sprays tend to be applied with herbicide applications either pre-seeding or before crop emergence, in order to control potential early-season pests (Gu et al. 2007; Micic et al. 2007; Lawrence 2009). Dependence on pesticides for insect control is not a new phenomenon. In 1959, VM Stern and colleagues (cited in Stern 1973) highlighted that insecticides were being applied without any understanding of the density at which pests caused economic injury to crops.

Thresholds that have been developed for pests of rainfed crops can be used to determine whether a pest needs to be controlled. If thresholds are used in combination with biological, cultural and chemical controls to decrease pest densities (i.e. integrated pest management, IPM), then pesticides are applied only when pest numbers exceed economic thresholds (Fick and Power 1992; Flint and Gouveia 2001; Flint et al. 2003; Edwards et al. 2008; Horne and Page 2008). However, the uptake and use of thresholds by farmers depends on the threshold being based on current research and taking into account the crop’s ability to tolerate damage. If the assessment methods needed to monitor for the threshold are too time-consuming, then prophylactic insecticide applications are more likely to occur (Gu et al. 2007; Leather and Atanasova 2017; Ramsden et al. 2017). Thresholds for some pests are shown in Table 3. In some cases, there are a number of thresholds for the same pest in the same crop, highlighting the need for a more coordinated approach for determining thresholds (e.g. Ramsden et al. 2017).


Table 3.  Some insect pests of rainfed crops and their control thresholds
Click to zoom

Integrated pest management

Reviews of IPM in Australia highlight its success in controlling pests in cotton and horticulture (see Williams and Il’ichev 2003; Gu et al. 2007; Horne et al. 2008; Hoy 2011). However, there are very limited examples of its success as a tactical management method in rainfed agriculture (Hoffmann et al. 2008). Horne et al. (2008) suggest that one of the factors responsible for poor adoption of IPM is that researchers have concentrated on a single pest and have not dealt with all of the pests in a crop. They also found that growers did not have the confidence to not apply a spray application for pests in crops such as canola and legumes.

For early-season pests (i.e. pests that affect crops at the seedling stage), the disadvantages of using a single ‘count’ threshold for a single pest can be overcome by applying an insecticide based on the amount of plant damage occurring to the crop at establishment (~14 days after germination) (e.g. Arthur et al. 2015). This approach relies on identifying pests that are present in a paddock from their feeding damage to the crop, and then applying an appropriate insecticide only if the crop is likely to fall below the optimum density for GY. For instance, for canola the target density should not be <20 plants/m2 (DPIRD 2019), for narrow-leaved lupin 35 plants/m2 (O’Connell et al. 2003), for barley 120/m2 (Paynter et al. 2019), and wheat 100 plants/m2 (Anderson et al. 2004) depending on target GY (see Table 1).

However, this approach requires use of a higher seeding rate to allow for some seedling loss. For instance, before hybrid canola seed was available, the recommendation was seeding at 5 kg/ha to establish ~50 seedlings/m2 (Micic 2005). However, the cost of hybrid canola seed is $34–48/ha, based on a seeding rate of 2 kg/ha (Seymour 2011), which leaves only 30 seedlings/m2. Optimal seedling density is 25 plants/m2 with conventional varieties and ≥30 seedlings/m2 with hybrid varieties (Seymour 2011). Consequently, a prophylactic spray of α-cypermethrin at 400 mL/ha, which costs ~$4.00/ha (Moore and Moore 2020), is an attractive option compared with reseeding a crop or increasing the seeding rate.

Tactical management for late-season pests such as diamond back moth (Plutella xylostella (L.)) and the armyworm complex (see Table 3) relies on monitoring crops until it is too late in the season to apply insecticides before harvest (Miller and Pike 2002; Floate and Hervet 2017). Crop monitoring can be time-intensive. The use of sweep nets to detect the presence of lepidopterous pests is a more time-efficient method than observing single plants. However, sweep nets are limited to the edges of crops such as canola owing to the difficulty of entering the crop at flowering (e.g. Floate and Hervet 2017). Sweep-net counts cannot be related to the density of the pest per plant (e.g. Dosdall et al. 2011); consequently, for the same pests there are different thresholds based on the monitoring technique used (see Table 3). The timing of crop monitoring needs to coincide with the presence of the pest in the landscape. This knowledge can be achieved through the use of pheromone traps or forecasting based on environment (e.g. Thackray et al. 2004; Harrington et al. 2007; Dosdall et al. 2011).

Timing of insecticide applications

Insecticides are usually applied when pests either cause sufficient plant loss or are at a threshold to cause yield loss. However, by understanding the lifecycle of pests, tactical control measures can be applied to suppress populations before egg laying occurs. For instance, the pest species most damaging to germinating canola in WA was the redlegged earth mite (Ridsdill-Smith et al. 2008). This mite hatches from over-summering eggs when there is sufficient moisture and when at least 7 days have elapsed with mean temperatures <20.5°C, with peak hatchings occurring in mid-May (Wallace 1970). This can coincide with the germination of seedling crops such as canola (Kirkegaard et al. 2016). One of the main recommendations for the suppression of redlegged earth mite for a pasture–canola cropping rotation is to apply control measures to pasture during spring. The timing of this spray will kill the female mites before over-summering eggs can be produced (Ridsdill-Smith et al. 2008) and leads to lower mite numbers in the following canola crop. However, in WA there has been a move from seeding canola in June to earlier in April (Glen 2000; Kirkegaard et al. 2016) so that crop germination is less likely to coincide with peak hatchings of redlegged earth mites.

However, early-autumn sowing means that temperatures can be milder and pests such as aphids (e.g. green peach aphid, Myzus persicae (Sulzer)) that are usually found in crops in spring are now present on seedling canola (Coutts et al. 2015; Macfadyn and Hill 2017). Aphids such as green peach aphid are resistant to many insecticides and can spread Turnip yellows virus, causing GY losses of 30% in canola crops if infected at the seeding stage (Coutts et al. 2015; de Little et al. 2017). Instead, the use of insecticide seed treatments is recommended to decrease the incidence of this aphid pest spreading virus to germinating crops (Coutts et al. 2015).

The use of insecticide seed treatments as a tactic to suppress pests has very low non-target effects and is compatible with an IPM framework (Williams 2017). However, insecticide seed dressings can have limitations in protecting seedlings from pest damage if pest numbers are high. For instance, seed dressings containing the active ingredient imidacloprid are only successful if mites such as redlegged earth mite are present in low densities (i.e. at thresholds of 10 mites/100 cm2) (Horne and Page 2008; Moore and Moore 2020).

Varietal selection

Selection of varieties that are resistant to pests and diseases is an important tactical management practice. Some crop cultivars have aphid resistance (e.g. Adhikari et al. 2012; Brewer et al. 2019); however, to date there are no commercialised varieties of canola, cereals or lupins with resistance to herbivorous invertebrates. There is some evidence that different varieties have different tolerances to damage (e.g. Liu and Ridsdill-Smith 2001). Even so, if crop varieties are chosen that match the predicted growing season and have good disease tolerance, crops are likely to be healthy (Williams 2017) and thus more likely to outgrow feeding damage, especially from establishment pests.

Implications for grain contamination

Invertebrate contaminants do not cause any injury to the grain but are incidentally harvested with the grain. There is a need to ensure high-quality grain for export; therefore, grain receival points have limits on the amount of contaminants. If this limit is exceeded, growers must arrange for their grain to be cleaned (Moore et al. 2019), an expense that is picked up directly by the grower. A tactical response to decrease contamination risk is to change time of harvest or harvest technique. For instance, if harvest occurs in the middle of the day when invertebrates are not actively moving there is less invertebrate contamination of grain; if crops are direct-harvested and not swathed, there is less contamination (Micic and Michael 2005; Micic et al. 2006).

Future implications for pest management

In dry conditions, farmers are more likely to dry-seed crops, with germination occurring as soon as there is sufficient soil moisture; this is a widespread practice in low-rainfall areas globally (ABARE 2017). Crops germinate with the first opening rains, which can coincide with warmer weather, leading to potentially moisture-stressed crops that are more susceptible to pest damage if insufficient follow-up rainfall occurs. Currently, the only tactical option if crop loss is occurring is to apply insecticides.

Longer term strategic management needs to focus on suppressing pest populations before they can cause crop damage regardless of the season. This can be achieved only by using cultural controls and requires forward planning because it cannot be applied in the year that economic damage is observed. Routine prophylactic spraying will increase the risk of insecticide resistance. This may be addressed by further research and extension on the economic effect of invertebrate pests on GYs and the efficacy of control measures.


Disease management

Management of the common diseases of the principal crops produced in winter-dominant rainfall regions often involves both tactical and strategic practices. Soil-, stubble- and air-borne diseases affect cereal crops grown in rainfed agricultural regions. The leaf diseases mostly require high humidity and either rainfall or dew on the leaves for germination of disease spores when the atmospheric temperatures are optimum, and wind for dispersal of spores (e.g. Jeger et al. 1981; Te Beest et al. 2008). Some examples are shown for barley (Table 4) and wheat (Table 5), indicating the specificity of the various diseases and the tactical management for them. Awareness of these factors is important for effective disease management tactics, but this may be further complicated by the different responses of some genotypes to environmental factors (Hogg et al. 1969; Garrett et al.2006).


Table 4.  Optimum conditions for infection of common leaf and stem diseases of barley and some methods used for tactical management
Click to zoom


Table 5.  Common leaf and stem diseases of wheat, their conditions required for disease development and some methods of tactical management.
Click to zoom

For stubble-borne diseases, the disease profile must be known. This entails knowledge of the diseases present in the paddock in the previous year and the susceptibility and tolerance of the crop that is to be grown in the current year. For air-borne diseases, the farmer also needs to know what diseases were prevalent in the previous season, the disease risks from any green bridge, and especially the diseases found upwind from the current crop.

Decision-support tools for disease management are limited and require development for necrotrophic pathogens such as septoria, net blotches and scalds, and biotrophic diseases such as powdery mildew. PREDICTA B is a support tool that is increasingly used for detection of soil-borne diseases (Ophel-Keller et al. 2008).

Variety selection

The use of resistant or tolerant varieties is the most economical way to manage disease in rainfed crops. Although growing resistant varieties does decrease disease risk, varieties also need to be high yielding for farmers to grow them. A single variety will not be resistant to all diseases present in the region; therefore, understanding the disease risk before sowing the crop is important for variety selection and disease management. Disease incidence varies with season, and variety susceptibility may also interact with seasonal influences. Variety selection can be facilitated through the yearly publication of variety sowing guides (e.g. Paynter et al. 2019; Zaicou-Kunesch et al. 2018) when the disease-resistance profile of varieties is known. However, this information is not always available and is a major research requirement in some regions.

The use of seed that has been certified free of relevant diseases, where available, can decrease seed-borne disease risk. For instance, every 1% of grain infected by loose smut leads to a 1% GY loss (Hills 2018). Those seeds need to be treated with fungicidal seed dressings to prevent further infection. As a minimum measure, farmer experience shows that knowing the provenance of imported seed is advisable if farm-saved seeds are not used.

Time of sowing

Sowing time can be altered to manage some crop diseases. For instance, in the southern parts of WA, spore showers of yellow spot (Pyrenophora tritici-repentis) occur in early April and May, and wheat crops sown at this time are most susceptible. If crops are sown later, disease risk is reduced. However, in northern areas of WA, yellow spot spores mature after the end of May and crops sown at this time are most at risk (Galloway et al. 2016).

Crop nutrition

Growing a healthy crop reduces some disease risk. For instance, incidence of barley powdery mildew and spot-type net blotch can be reduced by up to 40% in crops that have sufficient K. The combination of adequate K nutrition with foliar fungicides can further significantly reduce the impact of these diseases on GY (Brennan and Jayasena 2007). Conversely, application of high rates of N can promote the incidence of barley powdery mildew (Bainbridge 1974; Hills and Paynter 2007). However, late-season urea sprays can assist suppression of disease in winter wheat in some environments (Gooding et al. 1988).

Chemical control

If resistant varieties are not present for all common foliar diseases, fungicidal seed dressings can be used or fungicide can be applied in-furrow with the fertiliser at seeding. If seeds are not treated with fungicide, the seedling crop can be sprayed with a foliar fungicide. In modern cropping systems, farmers may focus on varieties with high GY irrespective of disease profile and may become heavily reliant on fungicide control. There is a need to consider an integrated approach to reduce reliance on fungicides so that these chemicals can remain effective in the future.

Implementing integrated control

Different diseases are controlled by different management practices. Examples of the components of integrated control measures for the major diseases of barley and wheat are summarised in Table 4 and 5 below.

Take-all (Gaeumannomyces spp.)

Take-all is a soil-borne root disease of wheat, barley, rye and oats, and is common in high-rainfall areas such as in the southern region of WA (MacLeod et al. 1993). It is caused by G. graminis vars tritici and avenae. Take-all is more severe in sandy, alkaline, infertile soils and exacerbated by continuous cropping, early sowing and poorly drained soils. The pathogen survives in summer in the cereal residues of the previous growing season’s grass hosts. The symptoms can be first noticed when premature plant death occurs during grain-filling, resulting in development of ‘white heads’. The affected plants often occur in conspicuous, circular patches in the field. Soil temperatures in the range 10–20°C are optimal for infection. Higher rainfall during winter favours early infection by the disease, but late infection near crop maturity has less impact on GY because infection is confined to roots.

Long-term continuous cereal cropping has been shown to decrease take-all over time (Kwak and Weller 2013; Lawes et al. 2013). Benefits arising from reduced take-all may be offset, however, by the economic losses arising from continuous cereal cropping (Loughman et al. 2000).

There is no varietal resistance available against take-all disease. Therefore, managing take-all relies primarily on cultural practices to prevent survival of the fungus between seasons. Practices such as eliminating grass hosts and use of non-cereal break crops (lupins, canola, field pea (Pisum sativum (L.)) in rotation can reduce carry-over of the disease between seasons. Minimising crop residues can also reduce take-all. However, no-till farming practices, which have been widely adopted to increase accumulation of organic matter and prevent soil erosion, serve to increase take-all. Burning the stubble can reduce the amount of infected surface residues but is not effective in eliminating the infected materials below the ground level.


Concluding remarks

Economic considerations

The ability to respond to weather conditions during the growing season may be complicated in the field by other limiting factors, especially soil physical problems. Farmers in both traditional and modern cropping systems may use conservative management practices and alter inputs in response to uncertain weather conditions and economic returns (e.g. Kingwell et al. 1993; Yigezu et al. 2014). However, when the causes for suboptimal performance are not fully understood, a process of diagnosis of the problems through local experimentation may well point the way to profitable GY improvement (Anderson et al. 2014).

The tactical management factors identified in this review all have a cost, or range of costs, and an anticipated benefit if effective. The impact on GY of each individual limiting factor can be estimated as a starting point for estimating the benefit and cost of remediating the factor. Comparisons of the range of benefits for the wheat crop from the tactical management factors can be made by examining published GY losses arising from individual limiting factors, along with some likely costs of treatment, possible benefits, and thus net benefits (Table 6). These costs are based on published values in the Australian literature, which in themselves show some variability. However, the remediated GYs in each of these publications were at, or close to, the estimated rainfall-limited potential GY in each case. In other regions where socioeconomic conditions may differ, the values listed in Table 6 should be used only to assess likely relative benefits of remediating the various limiting factors, rather than as absolute values.


Table 6.  Factors limiting grain yield of rainfed crops (wheat or barley), the likely losses, potential costs of remediation and possible net benefits.
Click to zoom

The data in Table 6 indicate that improving management of factors such as sowing time, nutrient application, or control of diseases, pests and weeds present at damaging levels can return far more to the farmer than changing the crop cultivar. This finding is similar to findings of Kirkegaard et al. (2014) in Australia and Piggin et al. (2015) in in northern Syria. However, many farmers prefer to adopt new crop cultivars rather than addressing other management factors, possibly because of the lower cost, logistical considerations such as availability of machinery, or not having the skills or confidence to implement other changes. The difficulty of resolving such tactical decisions in an environment where profit margins are small could be assisted by development of a simple decision aid (or crop model) based on the research discussed in this review. This decision aid could include an indication of the likely sequence of decisions and their linkage to seasonal conditions such as rainfall, temperature and humidity.

The issue of the potential overuse of chemical methods to manage the various biotic factors that can impact crop production cannot be ignored if sustained increases in productivity of rainfed systems are to be achieved. Increasing evidence of genetic resistance of various weed species to various chemicals (e.g. Owen and Powles 2018) is one particular challenge for researchers continuing to develop new, integrated systems that can be sustained into the future. The costs and ultimate benefits of developing integrated pest, weed and disease management systems requires further research, development and extension in order to achieve sustainable rainfed cropping systems (e.g. Nordblom 2003). The costs of remediation of these in-crop challenges relative to the potential benefits are quite small (see Table 6), so farmers tend to apply measures on the basis of very simple field observations rather than more detailed, randomised counts as might be appropriate for experiments.

Implications

Although the classical agronomic questions regarding practices such as sowing times, seed rates, and weed, disease and pest management have largely been addressed for the major cereal crops such as wheat and barley, widely applicable information is less available for the pulse crops and canola. The questions that remain are centred on integration of these practices with the longer term strategic issues largely involving soil improvement and crop rotations. Although we have confined our review to tactical management, there is a need for further examination of possible synergies between tactical, strategic and genetic methods in lifting average GYs. The relevance of research findings can be extended by using multi-site and multi-factor field experiments, for which there is a long tradition in Australia (e.g. French and Schultz 1984). Where it is desirable to add value to such data through development of crop models and decision aids, reliable field data can be invaluable for validation.

Many leading farmers have adopted the relevant research findings and their crop GYs are approaching the potential (e.g. Kingwell et al. 1993; Abeledo et al. 2008; Hochman et al. 2012; Robertson et al. 2012; Kirkegaard et al. 2014). Reasons for average or poorer GYs may be more related to operational convenience, availability of machinery, skills, and attitudes to economic risk rather than to maximising profits per se and they remain a challenge for researchers and advisers. Application of the findings regarding tactical management as outlined in this review can provide only a part of the solution to the problems that can be diagnosed on individual farms or soil types (e.g. Anderson et al. 2014). The onset of widespread climate change in the form of declining rainfall and changes in seasonal rainfall distribution (Stephens and Lyons 1998) is one of the main stimuli for reviewing some of the known means of addressing these changes.


Conflicts of interest

The authors declare no conflicts of interest.



Acknowledgements

Financial support by the Western Australian Government and the Grains Research and Development Corporation is acknowledged. Authors thank three anonymous reviewers and editor for helpful comments.


References

ABARE (2017) Australian crop report. Catalogue No. 183. Australian Bureau of Agricultural and Resource Economics, Canberra, ACT.

Abdul WK, Mann RA, Saleem M, Majeed A (2012) Comparative rice yield and economic advantage of foliar KNO3 over soil applied K2SO4. Pakistan Journal of Agricultural Sciences 49, 481–484.

Abeledo LG, Savin R, Slafer GA (2008) Wheat productivity in the Mediterranean Ebro Valley: analyzing the gap between attainable and potential yield with a simulation model. European Journal of Agronomy 28, 541–550.
Wheat productivity in the Mediterranean Ebro Valley: analyzing the gap between attainable and potential yield with a simulation model.Crossref | GoogleScholarGoogle Scholar |

Adhikari KN, Edwards OR, Wang S, Ridsdill-Smith JT, Buirchell B (2012) The role of alkaloids in conferring aphid resistance in yellow lupin (Lupinus luteus L.). Crop & Pasture Science 63, 444–451.
The role of alkaloids in conferring aphid resistance in yellow lupin (Lupinus luteus L.).Crossref | GoogleScholarGoogle Scholar |

Ali A, Hussain M, Saqib HH, Taj Kiani TT, Abbas Anees M, Abdul Rahman M (2016) Foliar spray surpasses soil application of potassium for maize production under rainfed conditions. Turkish Journal of Field Crops 21, 36–43.
Foliar spray surpasses soil application of potassium for maize production under rainfed conditions.Crossref | GoogleScholarGoogle Scholar |

Alrijabo AS (2014) Effect of a new farming system (zero-tillage) on the growth, yield and its components of bread wheat, durum wheat and barley crops in the moderate rainfall area of Ninevah Province. Mesopotamia Journal of Agriculture 42, 1–16.

Amede T, Tsegaye A (2016) Nurturing agricultural productivity and resilience in drylands of Sub-Saharan Africa. In ‘Innovations in dryland agriculture’. (Eds M Farooq, KHM Siddique) pp. 443–466. (Springer International Publishing: Cham, Switzerland)

Anderson WK (1985) Grain yield responses of barley and durum wheat to split nitrogen applications under rainfed conditions in a Mediterranean environment. Field Crops Research 12, 191–202.
Grain yield responses of barley and durum wheat to split nitrogen applications under rainfed conditions in a Mediterranean environment.Crossref | GoogleScholarGoogle Scholar |

Anderson WK (2010) Closing the gap between actual and potential yield of rainfed wheat. The impacts of environment, management and cultivar. Field Crops Research 116, 14–22.
Closing the gap between actual and potential yield of rainfed wheat. The impacts of environment, management and cultivar.Crossref | GoogleScholarGoogle Scholar |

Anderson WK, Sawkins D (1997) Production practices for improved grain yield and quality of soft wheats in Western Australia. Australian Journal of Experimental Agriculture 37, 173–180.
Production practices for improved grain yield and quality of soft wheats in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Anderson WK, Crosbie GB, Lemsom K (1995) Production practices for high protein, hard wheat in Western Australia. Australian Journal of Experimental Agriculture 35, 589–595.
Production practices for high protein, hard wheat in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Anderson WK, Heinrich A, Abbotts R (1996) Long season wheats extend sowing opportunities in the central wheat belt of Western Australia. Australian Journal of Experimental Agriculture 36, 203–208.
Long season wheats extend sowing opportunities in the central wheat belt of Western Australia.Crossref | GoogleScholarGoogle Scholar |

Anderson WK, Crosbie GB, Lambe WJ (1997) Production practices in Western Australia for wheats suitable for white, salted noodles. Australian Journal of Agricultural Research 48, 49–58.
Production practices in Western Australia for wheats suitable for white, salted noodles.Crossref | GoogleScholarGoogle Scholar |

Anderson WK, Sharma DL, Shackley BJ, D’Antuono MF (2004) Rainfall, sowing time, soil type and cultivar influence optimum plant population for wheat in Western Australia. Australian Journal of Agricultural Research 55, 921–930.
Rainfall, sowing time, soil type and cultivar influence optimum plant population for wheat in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Anderson WK, Hamza MA, Sharma DL, D’Antuono MF, Hoyle FC, Hill N, Shackley BJ, Amjad M, Zaicou-Kunesch C (2005) The role of management in yield improvement of the wheat crop—a review with special emphasis on Western Australia. Australian Journal of Agricultural Research 56, 1137–1149.
The role of management in yield improvement of the wheat crop—a review with special emphasis on Western Australia.Crossref | GoogleScholarGoogle Scholar |

Anderson WK, van Burgel AJ, Sharma DL, Shackley BJ, Zaicou-Kunesch CM, Miyan MS, Amjad M (2011) Assessing specific agronomic responses of wheat cultivars in a winter rainfall environment. Crop & Pasture Science 62, 115–124.
Assessing specific agronomic responses of wheat cultivars in a winter rainfall environment.Crossref | GoogleScholarGoogle Scholar |

Anderson WK, McTaggart RM, McQuade NC, Carter D, Overheu T, Bakker D, Peltzer S (2014) An approach to crop yield improvement through diagnostic systems research in a winter-dominant rainfall environment. Crop & Pasture Science 65, 922–933.
An approach to crop yield improvement through diagnostic systems research in a winter-dominant rainfall environment.Crossref | GoogleScholarGoogle Scholar |

Anderson G, Chen W, Bell RW, Brennan RF (2015) Making better fertiliser decisions for cropping systems in Western Australia. Soil test–crop response relationships and critical soil test values and ranges. Bulletin 4865. Department of Agriculture and Food Western Australia, Perth, W. Aust.

Anderson WK, Johansen C, Siddique KHM (2016) Addressing the yield gap in rainfed crops: a review. Agronomy for Sustainable Development 36, art.18.
Addressing the yield gap in rainfed crops: a review.Crossref | GoogleScholarGoogle Scholar |

Angus JF, Good JG (2004) Dryland cropping in Australia. In ‘Challenges and strategies for dryland agriculture’. (Eds SC Rao, J Ryan) pp. 151–166. (Crop Science Society of America: Madison, WI, USA)

Angus JF, van Heerwarden AF, Howe GN (1991) Productivity and break crop effects of winter-growing oilseeds. Australian Journal of Experimental Agriculture 31, 669–677.
Productivity and break crop effects of winter-growing oilseeds.Crossref | GoogleScholarGoogle Scholar |

Arabi MIE, Ali NM, Jahir M (2002) Effect of foliar and soil potassium fertilisation on yield and severity of Septoria tritici blotch. Australasian Plant Pathology 31, 359–362.
Effect of foliar and soil potassium fertilisation on yield and severity of Septoria tritici blotch.Crossref | GoogleScholarGoogle Scholar |

Arthur AL, Hoffmann AA, Umina PA (2015) Challenges in devising economic spray threshold for a major pest of Australian canola, the redlegged earth mite (Halotydeus destructor). Pest Management Science 71, 1462–1470.
Challenges in devising economic spray threshold for a major pest of Australian canola, the redlegged earth mite (Halotydeus destructor).Crossref | GoogleScholarGoogle Scholar | 25472683PubMed |

Australian Farm Institute (2012) Managing uncertainty in the world’s riskiest business. Farm Policy Journal 9, 1–72.

Ayesu-Offei EN, Carter MV (1971) Epidemiology of leaf scald of barley. Australian Journal of Agricultural Research 22, 383–390.
Epidemiology of leaf scald of barley.Crossref | GoogleScholarGoogle Scholar |

Bainbridge A (1974) Effect of nitrogen nutrition of the host on barley powdery mildew. Plant Pathology 23, 160–161.
Effect of nitrogen nutrition of the host on barley powdery mildew.Crossref | GoogleScholarGoogle Scholar |

Bakker DM, Hamilton DJ, Houlbrooke DJ, Span C, Van Burgel A (2007) Productivity of crops grown on raised beds on duplex soils prone to waterlogging in Western Australia. Australian Journal of Experimental Agriculture 47, 1368–1376.
Productivity of crops grown on raised beds on duplex soils prone to waterlogging in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Bastiaans L, Paolini R, Baumann DT (2008) Focus on ecological weed management: what is hindering adoption? Weed Research 48, 481–491.
Focus on ecological weed management: what is hindering adoption?Crossref | GoogleScholarGoogle Scholar |

Beard C, Thomas GJ (2018) Fungicides for managing powdery mildew in wheat historical trial report available at 1 May 2018. Available at: https://www.agric.wa.gov.au/grains-research-development/fungicides-managing-powdery-mildew-wheat-historical-trial-report (accessed 10 June 2020).

Beard C, Thomas G, Jayasena K (2018a) Managing stripe rust and leaf rust in wheat in Western Australia. Agriculture and Food, Department of Primary Industries and Regional Development, Government of Western Australia, Perth, W. Aust. Available at: https://www.agric.wa.gov.au/grains-research-development/managing-stripe-rust-and-leaf-rust-wheat-western-australia (accessed 23 May 2018).

Beard C, Jayasena K, Thomas GJ, Galloway J (2018b) Managing yellow spot and septoria nodorum blotch in wheat. Agriculture and Food, Department of Primary Industries and Regional Development, Government of Western Australia, Perth, W. Aust. Available at: https://www.agric.wa.gov.au/grains-research-development/managing-yellow-spot-and-septoria-nodorum-blotch-wheat?page=0%2C1 (accessed 13 September 2018).

Bell RW, Dell B (2008) ‘Micronutrients for sustainable food, feed, fibre and bio-energy production.’ (International Fertiliser Industry Association: Paris)

Bell LW, Moore AD, Kirkegaard JA (2014) Evolution in crop–livestock integration systems that improve farm productivity and environmental performance in Australia. European Journal of Agronomy 57, 10–20.
Evolution in crop–livestock integration systems that improve farm productivity and environmental performance in Australia.Crossref | GoogleScholarGoogle Scholar |

Bhathal JS, Loughman R, Speijers J (2003) Yield reduction in wheat in relation to leaf disease from yellow (tan) spot and Septoria nodorum blotch. European Journal of Plant Pathology 109, 435–443.
Yield reduction in wheat in relation to leaf disease from yellow (tan) spot and Septoria nodorum blotch.Crossref | GoogleScholarGoogle Scholar |

Bly AG, Woodard HJ (2003) Foliar nitrogen application timing on grain yield and protein concentration of hard red winter and spring wheat. Agronomy Journal 95, 335–338.
Foliar nitrogen application timing on grain yield and protein concentration of hard red winter and spring wheat.Crossref | GoogleScholarGoogle Scholar |

Bockus WW, Claassen MM (1992) Effects of crop rotation and residue management practices on severity of tan spot of winter wheat. Plant Disease 76, 633–636.
Effects of crop rotation and residue management practices on severity of tan spot of winter wheat.Crossref | GoogleScholarGoogle Scholar |

Brennan RF (1989) Effect of superphosphate and superphosphate plus flutriafol on yield and take-all of wheat. Australian Journal of Experimental Agriculture 29, 247–252.
Effect of superphosphate and superphosphate plus flutriafol on yield and take-all of wheat.Crossref | GoogleScholarGoogle Scholar |

Brennan RF (1991) Effectiveness of zinc sulphate and zinc chelate as foliar sprays for alleviating zinc deficient soils in Western Australia. Australian Journal of Experimental Agriculture 31, 831–834.
Effectiveness of zinc sulphate and zinc chelate as foliar sprays for alleviating zinc deficient soils in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Brennan RF (2000) Copper, zinc, molybdenum and boron. In ‘The wheat book—principles and practice’. Bulletin 4443. (Eds WK Anderson, JR Garlinge) pp. 95–105. (Department of Agriculture Western Australia: Perth, W. Aust.)

Brennan RF (2006) Residual value of molybdenum for wheat production on naturally acidic soils of Western Australia. Australian Journal of Experimental Agriculture 46, 1333–1339.
Residual value of molybdenum for wheat production on naturally acidic soils of Western Australia.Crossref | GoogleScholarGoogle Scholar |

Brennan RF, Bolland MDA (2008) Significant nitrogen by sulfur interactions occurred for canola grain production and oil concentration in grain on sandy soils in the Mediterranean-type climate of south Western Australia. Journal of Plant Nutrition 31, 1174–1187.
Significant nitrogen by sulfur interactions occurred for canola grain production and oil concentration in grain on sandy soils in the Mediterranean-type climate of south Western Australia.Crossref | GoogleScholarGoogle Scholar |

Brennan RF, Bolland MDA (2011) Foliar spray experiments identify no boron deficiency but molybdenum and manganese deficiency for canola grain production on acidified sandy gravel soils in south Western Australia. Journal of Plant Nutrition 34, 1186–1197.
Foliar spray experiments identify no boron deficiency but molybdenum and manganese deficiency for canola grain production on acidified sandy gravel soils in south Western Australia.Crossref | GoogleScholarGoogle Scholar |

Brennan RF, Jayasena KW (2007) Increasing applications of potassium fertiliser to barley crops grown on deficient sandy soils increased grain yields while decreasing some foliar diseases. Australian Journal of Agricultural Research 58, 680–689.
Increasing applications of potassium fertiliser to barley crops grown on deficient sandy soils increased grain yields while decreasing some foliar diseases.Crossref | GoogleScholarGoogle Scholar |

Brennan RF, Bell RW, Frost K (2015) Risks of boron toxicity in canola and lupin by forms of boron application in acid sands of South-Western Australia. Journal of Plant Nutrition 38, 920–937.
Risks of boron toxicity in canola and lupin by forms of boron application in acid sands of South-Western Australia.Crossref | GoogleScholarGoogle Scholar |

Brewer MJ, Peairs FB, Elliott NC (2019) Invasive cereal aphids of North America: ecology and pest management. Annual Review of Entomology 64, 73–93.
Invasive cereal aphids of North America: ecology and pest management.Crossref | GoogleScholarGoogle Scholar | 30372159PubMed |

Brown MP, Steffenson B, Webster RK (1993) Host range of Pyrenophora teres f. teres isolates from California. Plant Disease 77, 942–947.
Host range of Pyrenophora teres f. teres isolates from California.Crossref | GoogleScholarGoogle Scholar |

Burdon JJB, Abbott DC, Brown AHD, Brown JS (1994) Genetic structure of the Scald pathogen (Rhynchosporium secalis). Australian Journal of Agriculture Research 45, 1445–1454.
Genetic structure of the Scald pathogen (Rhynchosporium secalis).Crossref | GoogleScholarGoogle Scholar |

Canola Council of Canada (2014) Diamondback moth. Canola Encyclopedia. Canola Council of Canada, Winnipeg, MB, Canada. Available at: https://www.canolacouncil.org/canola-encyclopedia/insects/diamondback-moth/#fnote26

Cargill (2020a) AWB daily contract prices. Bid sheet for wheat, Western Australia. Cargill Australia, Melbourne, Vic. Available at: http://image.info.cargill.com/lib/fe911574736c0c7e75/m/1/Wheat_WA.pdf (accessed 10 June 2020).

Cargill (2020b) AWB daily contract prices. Bid sheet for barley, Western Australia. Cargill Australia, Melbourne, Vic. Available at: http://image.info.cargill.com/lib/fe911574736c0c7e75/m/1/Barley_Feed_WA.pdf (accessed 10 June 2020).

Cargill (2020c) AWB daily contract prices. Bid sheet for canola, South Australia and Western Australia. Cargill Australia, Melbourne, Vic. Available at: http://image.info.cargill.com/lib/fe911574736c0c7e75/m/1/Oilseeds_WA_SA.pdf (accessed 10 June 2020).

Carter DJ, Gilkes RJ, Walker F (1998) Claying of water repellent soils: Effects on hydrophobicity, organic matter and nutrient uptake. In ‘Proceedings 16th World Congress of Soil Science’. Montpellier, France. (International Union of Soil Science)

Christensen NW, Mients VW (1982) Evaluating N sources and timing for winter wheat. Agronomy Journal 74, 840–844.
Evaluating N sources and timing for winter wheat.Crossref | GoogleScholarGoogle Scholar |

Christiansen S, Ryan J, Singh M, Ates S, Bahhady F, Mohamed K, Youssef O, Loss S (2015) Potential legume alternatives to fallow and wheat monoculture for Mediterranean environments. Crop & Pasture Science 66, 113–121.
Potential legume alternatives to fallow and wheat monoculture for Mediterranean environments.Crossref | GoogleScholarGoogle Scholar |

Cooper JE (1974) Effects of post-planting applications of nitrogenous fertilisers on grain yield, grain protein content and mottling of wheat. Queensland Journal of Agricultural and Animal Sciences 31, 33–42.

Coutts BA, Jones RAC, Umina P, Davidson J, Baker G, Aftab M (2015) Beet western yellows virus (synonym: Turnip yellows virus) and green peach aphid in canola. GRDC Update Papers, 10 Feb. 2015. Grains Research and Development Corporation, Canberra, ACT. Available at: https://grdc.com.au/resources-and-publications/grdc-update-papers/tab-content/grdc-update-papers/2015/02/beet-western-yellows-virus-synonym-turnip-yellows-virus-and-green-peach-aphid-in-canola

Coutts B, Thomas GJ, Loughman R, Huberli D, Micic S (2018) Control of green bridge for pest and disease management. Department of Agriculture and Food Western Australia, Perth, W. Aust. Available at: https://www.agric.wa.gov.au/grains/control-green-bridge-pest-and-disease-management?page=0%2C0 (accessed 10 June 2020).

Cunfer BM, Touchton JT, Johnson JW (1980) Effects of phosphorous and potassium fertilisation on a Septoria glume blotch of wheat. Phytopathology 70, 1196–1199.
Effects of phosphorous and potassium fertilisation on a Septoria glume blotch of wheat.Crossref | GoogleScholarGoogle Scholar |

D’Emden F, Llewellyn RS (2006) No-till adoption decisions in southern Australian cropping and the role of weed management. Australian Journal of Experimental Agriculture 46, 563–569.

de Little SC, Edwards O, van Rooyen AR, Weeks A, Umina PA (2017) Discovery of metabolic resistance to neonictinoids in green peach aphid (Myzus persicae) in Australia. Pest Management Science 73, 1611–1617.
Discovery of metabolic resistance to neonictinoids in green peach aphid (Myzus persicae) in Australia.Crossref | GoogleScholarGoogle Scholar | 27888606PubMed |

Del Cima R, D’Antuono MF, Anderson WK (2004) The effects of soil type and seasonal rainfall on the optimum seed rate for wheat in Western Australia. Australian Journal of Experimental Agriculture 44, 585–594.
The effects of soil type and seasonal rainfall on the optimum seed rate for wheat in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Dell B, Huang L, Bell RW (2002) Boron in plant reproduction. In ‘Boron in plants and animal nutrition’. (Eds HE Goldbach, B Rerkasem M Wimmer, PH Brown, M Thellier, RW Bell) pp. 103–118. (Kluwer Academic/Plenum Publishers: New York) doi.org/10.1007/978-1-4615-0607-2

Delserone LM, Cole H (1987) Effect of planting date on development of net blotch epidemics in winter barley in Pennsylvania. Plant Disease 71, 438–441.
Effect of planting date on development of net blotch epidemics in winter barley in Pennsylvania.Crossref | GoogleScholarGoogle Scholar |

Derpsch R (2003) Conservation tillage, no-tillage and related technologies. In ‘Conservation agriculture: environment, farmer experiences, innovations, social economy, policy’. (Eds L Garcia-Torres, J Benites, A Martinez-Vilela, A Holgado-Cabrera) pp. 183–190. (Kluwer Academic Publishers: Dordrecht, The Netherlands)

Dewar AM (2017) Chemical control. In ‘Aphids as crop pests’. 2nd edn. (Eds HF Van Emden, R Harrington) pp. 391–423. (CABI Publishing: Wallingford, UK)

Dodd J, Martin RJ, Howes KM (1993) ‘Management of agricultural weeds in Western Australia.’ Bulletin 4243. (Western Australian Department of Agriculture: Perth, W. Aust.)

Donald CM (1962) In search of yield. Journal of the Australian Institute of Agricultural Science 28, 171–178.

Doole G, Weetman E (2009) Tactical management of pasture fallows in Western Australian cropping systems. Agricultural Systems 102, 24–32.
Tactical management of pasture fallows in Western Australian cropping systems.Crossref | GoogleScholarGoogle Scholar |

Dosdall LM, Soroka JJ, Olfert O (2011) The diamondback moth in canola and mustard: current pest status and future pest prospects. Prairie Soils and Crops Journal 4, 66–76.

Doyle AD, Marcellos H (1974) Time of sowing and wheat yield in northern New South Wales. Australian Journal of Experimental Agriculture and Animal Husbandry 14, 93–102.
Time of sowing and wheat yield in northern New South Wales.Crossref | GoogleScholarGoogle Scholar |

Doyle RJ, Parkin RJ, Smith AC, Gartrell JW (1965) Molybdenum increases cereal yields on wheatbelt sandplain. Journal of the Department of Agriculture Western Australia 66, 699–703.

DPIRD (2019) Information for the canola seeding rate calculator. Department of Primary Industries and Regional Development. Available at: www.agric.wa.gov.au/canola/information-canola-seeding-rate-calculator (accessed 10 June 2020).

Duczek LJ, Sutherland KA, Reed SL, Bailey KL, Lafond GP (1999) Survival of leaf spot pathogens on crop residues of wheat and barley in Saskatchewan. Canadian Journal of Plant Pathology 21, 165–173.
Survival of leaf spot pathogens on crop residues of wheat and barley in Saskatchewan.Crossref | GoogleScholarGoogle Scholar |

Edwards O, Franzmann B, Thackray D, Micic S (2008) Insecticide resistance and implications for future aphid management in Australian grains and pastures: a review. Australian Journal of Experimental Agriculture 48, 1523–1530.
Insecticide resistance and implications for future aphid management in Australian grains and pastures: a review.Crossref | GoogleScholarGoogle Scholar |

Ellen J, Spiertz JHJ (1980) Effects of rate and timing of nitrogen dressings on grain yield formation of winter wheat (T. aestivum L.). Fertilizer Research 1, 177–190.
Effects of rate and timing of nitrogen dressings on grain yield formation of winter wheat (T. aestivum L.).Crossref | GoogleScholarGoogle Scholar |

Eyal Z (1999) The septoria tritici and stagonospora nodorum blotch diseases of wheat. European Journal of Plant Pathology 105, 629–641.
The septoria tritici and stagonospora nodorum blotch diseases of wheat.Crossref | GoogleScholarGoogle Scholar |

Fageria NK (2009) ‘The use of nutrients in crop plants.’ (CRC Press: Boca Raton, FL, USA)

Fageria NK, Barbosa Filho MP, Moreira A, Guimarães CM (2009) Foliar fertilisation of crop plants. Journal of Plant Nutrition 32, 1044–1064.
Foliar fertilisation of crop plants.Crossref | GoogleScholarGoogle Scholar |

Farooq M, Flower KC, Jabran K, Wahid A, Siddique KHM (2011) Crop yield and weed management in rainfed conservation agriculture. Soil & Tillage Research 117, 172–183.
Crop yield and weed management in rainfed conservation agriculture.Crossref | GoogleScholarGoogle Scholar |

Farooqi MQU, Ahmad R, Wariach EA, Arfan M (2012) Effect of supplemental foliar applied potassium on yield and grain quality of autumn planted maize (Zea mays L.) under water stress. Journal of Food Agriculture and Veterinary Science 2, 8–12.

Fernández V, Brown PH (2013) From plant surface to plant metabolism: the uncertain fate of foliar-applied nutrients. Frontiers in Plant Science 4, 1–15.
From plant surface to plant metabolism: the uncertain fate of foliar-applied nutrients.Crossref | GoogleScholarGoogle Scholar |

Fernández V, Guzmán P, Peirce CAE, McBeath TM, Khayet M, McLaughlin MJ (2014) Effect of wheat phosphorus status on leaf surface properties and permeability to foliar-applied phosphorus. Plant and Soil 384, 7–20.
Effect of wheat phosphorus status on leaf surface properties and permeability to foliar-applied phosphorus.Crossref | GoogleScholarGoogle Scholar |

Fick GW, Power AG (1992) Pests and integrated control. In ‘Ecosystems of the world: field crop ecosystems’. (Ed. CJ Pearson) pp. 59–83. (Elsevier Science: Amsterdam)

Finney KF, Meyer JW, Smith FW, Fryer HC (1957) Effect of foliar spraying of Pawnee wheat with urea solutions on yield, protein content and protein quality. Agronomy Journal 49, 341–347.
Effect of foliar spraying of Pawnee wheat with urea solutions on yield, protein content and protein quality.Crossref | GoogleScholarGoogle Scholar |

Fischer RA, Edmeades GO (2010) Breeding and cereal yield progress. Crop Science 50, 585–598.
Breeding and cereal yield progress.Crossref | GoogleScholarGoogle Scholar |

Fischer RA, Byerlee D, Edmeades GO (2014) Crop yields and global food security: will yield increase continue to feed the world? Australian Centre for International Agricultural Research, Canberra, ACT. Available at: http://aciar.gov.au/publication/mn158

Fletcher A, Ogden G, Sharma D (2019) Mixing it up – wheat cultivar mixtures can increase yield and buffer the risks of flowering too early or too late. European Journal of Agronomy 103, 90–97.
Mixing it up – wheat cultivar mixtures can increase yield and buffer the risks of flowering too early or too late.Crossref | GoogleScholarGoogle Scholar |

Flint ML, Gouveia P (2001) ‘IPM in practice: principles and methods of integrated pest management.’ Publication 3418. (University of California: Oakland, CA, USA)

Flint ML, Daar S, Milinar R (2003) ‘Establishing integrated pest management policies and programs: a guide for public agencies.’ Publication 8093. (University of California: Oakland, CA, USA)

Floate KD, Hervet VA (2017) Noctuid (Lepidoptera: Noctuidae) pests of canola in North America. In ‘Integrated management of insect pests on canola and other brassica oilseed crops’. (Ed. GVP Reddy) (CABI Publishing: Wallingford, UK)

Francki MG, Shankar M, Walker E, Loughman R, Golzar H, Ohm H (2011) New quantitative trait loci in wheat for flag leaf resistance to Stagonospora nodorum blotch. Phytopathology 101, 1278–1284.
New quantitative trait loci in wheat for flag leaf resistance to Stagonospora nodorum blotch.Crossref | GoogleScholarGoogle Scholar | 21770777PubMed |

Franke W (1967) Mechanisms of foliar penetration of solutions. Annual Review of Plant Physiology 18, 281–300.
Mechanisms of foliar penetration of solutions.Crossref | GoogleScholarGoogle Scholar |

Franzluebbers A, Steiner J, Karlen D, Griffin T, Singer J, Tanaka D (2011) Rainfed farming systems in the USA. In ‘Rainfed farming systems’. (Eds P Tow, I Cooper, I Partridge, C Birch) pp. 511–560. (Springer: Dordrecht, The Netherlands)

French RJ, Schultz JE (1984) Water use efficiency of wheat in a Mediterranean-type environment. II. Some limitations to efficiency. Australian Journal of Agricultural Research 35, 765–775.
Water use efficiency of wheat in a Mediterranean-type environment. II. Some limitations to efficiency.Crossref | GoogleScholarGoogle Scholar |

French RJ, McCarthy K, Smart WL (1994) Optimum plant population densities for lupin (Lupinus angustifolius L.) in the wheatbelt of Western Australia. Australian Journal of Experimental Agriculture 34, 491–497.
Optimum plant population densities for lupin (Lupinus angustifolius L.) in the wheatbelt of Western Australia.Crossref | GoogleScholarGoogle Scholar |

French RJ, Seymour M, Malik RS (2016) Plant density response and optimum crop densities for canola (Brassica napus L.) in Western Australia. Crop & Pasture Science 67, 397–408.
Plant density response and optimum crop densities for canola (Brassica napus L.) in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Galloway J, Salam M, Payne P (2016) Less yellow spot in wheat: progress towards a decision support tool that will predict when yellow spot spores are released from the previous season stubble. 2016 GRDC Grains Research Updates. GIWA, Perth, W. Aust. Available at: http://www.giwa.org.au/2016researchupdates

Garrett KA, Dendy SP, Frank EE, Rouse MN, Travers SE (2006) Climate change effects on plant disease: genomes to ecosystems. Annual Review of Phytopathology 44, 489–509.
Climate change effects on plant disease: genomes to ecosystems.Crossref | GoogleScholarGoogle Scholar | 16722808PubMed |

Gartrell JW (1981) Distribution and correction of Cu deficiency in crops and pastures. In ‘Copper in soils and plants’. (Eds JF Loneragan, AD Robson, RD Graham) pp. 313–330. (Academic Press: Sydney)

Gartrell JW, Bolland MDA (2000) Nutrient removal in wheat crop products. In ‘The wheat book–principles and practice’. (Eds WK Anderson, JR Garlinge) pp. 106–107. (Department of Agriculture Western Australia: Perth, W. Aust.)

Gettier SW, Martens DC, Brumback TB (1985) Timing of foliar manganese application for correction of manganese deficiency in soybeans. Agronomy Journal 77, 627–630.
Timing of foliar manganese application for correction of manganese deficiency in soybeans.Crossref | GoogleScholarGoogle Scholar |

Gigot C, Saint-Jean S, Huber L, Maumene C, Leconte M, Kerhornou B, de Vallavieille-Pope C (2013) Protective effects of a wheat cultivar mixture against splash-dispersed septoria tritici blotch epidemics. Plant Pathology 62, 1011–1019.
Protective effects of a wheat cultivar mixture against splash-dispersed septoria tritici blotch epidemics.Crossref | GoogleScholarGoogle Scholar |

Glen DM (2000) The effects of cultural measures on cereal pests and their role in integrated pest management. Integrated Pest Management Reviews 5, 25–40.
The effects of cultural measures on cereal pests and their role in integrated pest management.Crossref | GoogleScholarGoogle Scholar |

Godfray J, Charles JR, Beddington IR, Crute L, Haddad D, Lawrence JF, Muir J, Pretty S, Robinson S, Thomas M, Toulmin C (2010) Food security: the challenge of feeding 9 billion people. Science 327, 812–818.
Food security: the challenge of feeding 9 billion people.Crossref | GoogleScholarGoogle Scholar |

Golzar H, Shankar M, D’Antuono M (2016) Responses of commercial wheat varieties and differential lines to Western Australian powdery mildew (Blumeria graminis f. sp. tritici) populations. Australasian Plant Pathology 45, 347–355.
Responses of commercial wheat varieties and differential lines to Western Australian powdery mildew (Blumeria graminis f. sp. tritici) populations.Crossref | GoogleScholarGoogle Scholar |

Gooding MJ, Davies WP (1992) Foliar urea fertilisation of cereals: a review. Fertilizer Research 32, 209–222.
Foliar urea fertilisation of cereals: a review.Crossref | GoogleScholarGoogle Scholar |

Gooding MJ, Kettlewell PS, Davies WP (1988) Disease suppression by late season urea sprays on winter wheat and interaction with fungicide. Journal of Fertilizer Issues 5, 19–23.

Graham RD (2008) Micronutrient deficiencies in crops and their global significance. In ‘Micronutrient deficiencies in global crop production’. (Ed. BJ Alloway) pp. 41–61. (Springer: Dordrecht, The Netherlands)

Grain Central (2019) Lupin crop seen at 405,700 t: Pulse Australia. 26 November 2019. Grain Central, Brisbane, Qld. Available at: https://www.graincentral.com/markets/lupin-crop-seen-at-482700t-pulse-australia/

Grimm M (1995) Insect pests of barley. In ‘The barley book’. (Ed. KL Young) pp. 63–69. (Department of Agriculture Western Australia: Perth, W. Aust.)

Grundon NJ (1980) Effectiveness of soil dressings and foliar sprays of Cu sulphate in correcting Cu deficiency of wheat (Triticum aestivum) in Queensland. Australian Journal of Experimental Agriculture and Animal Husbandry 20, 717–723.
Effectiveness of soil dressings and foliar sprays of Cu sulphate in correcting Cu deficiency of wheat (Triticum aestivum) in Queensland.Crossref | GoogleScholarGoogle Scholar |

Grundy AC, Mead A (1998) Modelling the effects of seed depth on weed seedling emergence weed seedbanks: determination, dynamics and manipulation. Aspects of Applied Biology 51, 75–82.

Gu H, Fitt P, Baker GH (2007) Invertebrate pests of canola and their management in Australia: a review. Australian Journal of Entomology 46, 231–243.
Invertebrate pests of canola and their management in Australia: a review.Crossref | GoogleScholarGoogle Scholar |

Guler M, Karaca M (1988) Agroclimatological criteria for determining the boundaries of fallow practice. In ‘Winter cereals and food legumes in mountainous areas’. (Eds JP Srivastava, C Saxena, S Varma, M Tahir) pp. 41–49. (International Center for Agricultural Research in the Dry Areas: Aleppo, Syria)

Hamza MA, Anderson WK (2003) Responses of soil properties and grain yield to deep ripping and gypsum application in a compacted loamy sand soil contrasted with a sandy clay loam soil in Western Australia. Australian Journal of Agricultural Research 54, 273–282.
Responses of soil properties and grain yield to deep ripping and gypsum application in a compacted loamy sand soil contrasted with a sandy clay loam soil in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Hamza MA, Anderson WK (2005) Soil compaction in cropping systems. A review of the nature, causes and possible solutions. Soil and Tillage Research 82, 121–145.

Harrington R, Hulle M, Plantegenest M (2007) Monitoring and forecasting. In ‘Aphids as crop pests’. (Eds H van Emden, R Harrington) pp. 515–536. (CABI Publishing: Wallingford, UK)

Hess DE, Shaner G (1987a) Effect of moisture and temperature on development of Septoria tritici blotch in wheat. Phytopathology 77, 215–219.
Effect of moisture and temperature on development of Septoria tritici blotch in wheat.Crossref | GoogleScholarGoogle Scholar |

Hess DE, Shaner G (1987b) Effect of moisture on Septoria tritici blotch development on wheat in the field. Phytopathology 77, 220–226.
Effect of moisture on Septoria tritici blotch development on wheat in the field.Crossref | GoogleScholarGoogle Scholar |

Hills A (2018) Controlling barley loose smut. Department of Agriculture and Food Western Australia, Perth, W. Aust. Available at: Available at: https://www.agric.wa.gov.au/barley/controlling-barley-loose-smut-2015

Hills A, Jayasena KW (2013) Spiroxamine offers an additional cost effective mode of action against fungicide resistant powdery mildew. In ‘Proceedings 2013 Agribusiness Crop Updates’. (GIWA: Perth, W. Aust.) Available at: http://www.giwa.org.au/2013-crop-updates

Hills A, Paynter B (2007) Barley agronomy highlights: canopy management. In ‘Proceedings 2007 Agribusiness Crop Updates’. (Department of Agriculture and Food Western Australia: Perth, W. Aust.) Available at: https://www.researchgate.net/publication/273135394

Hills A, Paynter B (2012) Grazing barley controls early foliar diseases, has manageable impacts on malting barley grain quality but suffers a yield penalty. In ‘Capturing opportunities and overcoming obstacles in Australian Agronomy. Proceedings 16th Australian Agronomy Conference’. University of New England, Armidale, NSW. (Australian Society of Agronomy) Available at: http://agronomyaustraliaproceedings.org/images/sampledata/2012/7915_8_hills.pdf

Hochman Z, Gobbett D, Holzworth D, McClelland T, van Rees H, Marinoni O, Navarro Garcia J, Horan H (2012) Quantifying yield gaps in rainfed cropping systems: a case study of wheat in Australia. Field Crops Research 136, 85–96.
Quantifying yield gaps in rainfed cropping systems: a case study of wheat in Australia.Crossref | GoogleScholarGoogle Scholar |

Hoffmann AA, Weeks AR, Nash MA, Mangano GP, Umina PA (2008) The changing status of invertebrate pests and the future of pest management in the Australian grains industry. Australian Journal of Experimental Agriculture 48, 1481–1493.
The changing status of invertebrate pests and the future of pest management in the Australian grains industry.Crossref | GoogleScholarGoogle Scholar |

Hogg WH, Hounam CE, Mallik AK, Zodoks JC (1969) Meteorological factors affecting the epidemiology of wheat rusts. Working Group on Meteorological Factors Affecting the Epidemiology of Wheat Rusts of the Commission for Agricultural Meteorology. Technical Note 99. World Meteorological Organisation, Geneva.

Horne PA, Page J (2008) ‘Integrated pest management for crops and pastures.’ (Landlinks Press: Melbourne, Vic.)

Horne PA, Page J, Nicholson C (2008) When will integrated pest management strategies be adopted? Example of the development and implementation of integrated pest management strategies in cropping systems in Victoria. Australian Journal of Experimental Agriculture 48, 1601–1607.
When will integrated pest management strategies be adopted? Example of the development and implementation of integrated pest management strategies in cropping systems in Victoria.Crossref | GoogleScholarGoogle Scholar |

Hoy M (2011) ‘Agricultural acarology: an introduction to integrated mite management.’ (CRC Press: Boca Raton, FL, USA)

James DG (2000) Abundance and phenology of earth mites (Acari: Penthaleidae) and predatory mites in pesticide treated and pesticide-free grassland habitats in southern New South Wales, Australia. International Journal of Acarology 26, 363–369.
Abundance and phenology of earth mites (Acari: Penthaleidae) and predatory mites in pesticide treated and pesticide-free grassland habitats in southern New South Wales, Australia.Crossref | GoogleScholarGoogle Scholar |

Jarvis RJ, Bolland MDA (1990) Placement of fertilisers for cereal crops. Fertilizer Research 22, 97–107.
Placement of fertilisers for cereal crops.Crossref | GoogleScholarGoogle Scholar |

Jayasena KW, Loughman R (2001) Integrated management of spot-type net blotch in barley. In ‘Proceedings Australasian Plant Pathology Society 13th Biennial Conference’. 24–27 September 2001. (Eds V Oliver, P Trevorrow, R Davis) p. 159. (Queensland Department of Primary Industries: Brisbane, Qld) https://trove.nla.gov.au/version/46557517 (accessed 10 June 2020)

Jayasena KW, Loughman R, Tanaka K (2006) Late-season management of powdery mildew in barley with foliar fungicides. Australasian Plant Pathology 35, 355–357.
Late-season management of powdery mildew in barley with foliar fungicides.Crossref | GoogleScholarGoogle Scholar |

Jayasena KW, Thomas G, Gupta S, Hills A, Wahlsten L (2018) Where we are with Adult Plant Resistance varieties for barley leaf rust - is it worth spraying? 2018 GRDC Grains Research Updates. GIWA, Perth, W. Aust. Available at: http://www.giwa.org.au/2018researchupdates

Jeger MJ, Griffiths E, Jones DG (1981) Influence of environmental conditions on spore dispersal and infection by Septoria nodorum. Annals of Applied Biology 99, 29–34.
Influence of environmental conditions on spore dispersal and infection by Septoria nodorum.Crossref | GoogleScholarGoogle Scholar |

Jenkins JEE, Jemmett JL (1967) Barley leaf blotch. N.A.A.S. The Quarterly Review 75, 127–132.

Jettner R, Loss SP, Martin JD, Siddique KHM (1998) Responses of faba bean (Vicia faba L.) to sowing rate in south-western Australia. 1. Seed yield and economic optimum plant density. Australian Journal of Agricultural Research 49, 989–998.
Responses of faba bean (Vicia faba L.) to sowing rate in south-western Australia. 1. Seed yield and economic optimum plant density.Crossref | GoogleScholarGoogle Scholar |

John LJ, Lester GE (2011) Effect of foliar potassium fertilization and source on cantaloupe yield and quality. Better Crops with Plant Food 95, 13–15.

Johnston RL, Bishop GW (1987) Economic injury levels and economic thresholds for cereal aphids (Homoptera: Aphididae) on spring-planted wheat. Journal of Economic Entomology 80, 478–482.
Economic injury levels and economic thresholds for cereal aphids (Homoptera: Aphididae) on spring-planted wheat.Crossref | GoogleScholarGoogle Scholar |

Jones RA (1994) Effect of mulching with cereal straw and row spacing on spread of bean yellow mosaic potyvirus into narrow-leafed lupins (Lupinus angustifolius). Annals of Applied Biology 124, 45–58.
Effect of mulching with cereal straw and row spacing on spread of bean yellow mosaic potyvirus into narrow-leafed lupins (Lupinus angustifolius).Crossref | GoogleScholarGoogle Scholar |

Jones RE, Vere DT, Alemseged Y, Medd RW (2005) Estimating the economic cost of weeds in Australian annual winter crops. Agricultural Economics 32, 253–265.
Estimating the economic cost of weeds in Australian annual winter crops.Crossref | GoogleScholarGoogle Scholar |

Kearns S, Umbers A (2010) 2010 GRDC farm practice baseline report. Grains Research and Development Corporation, Canberra, ACT. Available at: https://www.grdc.com.au/__data/assets/pdf_file/0022/209740/grdc-farm-practices-baseline-full-report-july-2010.pdf

Keating BA, Carberry PS, Hammer GL, Probert ME, Robertson MJ, Holzworth D, Huth NI, Hargreaves JNG, Meinke H, Hochman Z, McLean G, Verburg K, Snow V, Dimes JP, Silburn M, Wang E, Brown S, Bristow KL, Asseng S, Chapman S, McCown RL, Freebairn DM, Smith CJ (2003) An overview of APSIM, a model designed for farming systems simulation. European Journal of Agronomy 18, 267–288.
An overview of APSIM, a model designed for farming systems simulation.Crossref | GoogleScholarGoogle Scholar |

Khan TN (1986) Effects of fungicide treatments on scald (Rhynchosporium secalis) infection and yield of barley in Western Australia. Australian Journal of Experimental Agriculture 26, 231–235.
Effects of fungicide treatments on scald (Rhynchosporium secalis) infection and yield of barley in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Khan TN (1988) Effects of stubble-borne fungal inoculum on incidence of leaf diseases and yields of barley in Western Australia. Australian Journal of Experimental Agriculture 28, 529–532.
Effects of stubble-borne fungal inoculum on incidence of leaf diseases and yields of barley in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Khan TN (1989) Effect of spot-type net blotch (Drechslera teres (Sacc.) Schoem) infection on barley yield in short season environment of northern cereal belt of Western Australia. Australian Journal of Agricultural Research 40, 745–752.
Effect of spot-type net blotch (Drechslera teres (Sacc.) Schoem) infection on barley yield in short season environment of northern cereal belt of Western Australia.Crossref | GoogleScholarGoogle Scholar |

Khan TN, D’Antuono MF (1985) Relationship between scald (Rhynchosporium secalis) and losses in grain yield of barley in Western Australia. Australian Journal of Agricultural Research 36, 655–661.
Relationship between scald (Rhynchosporium secalis) and losses in grain yield of barley in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Khan TN, Young KJ (1988) Seed dressing with Baytan and barley yield in Western Australia. Australasian Plant Pathology 17, 99–100.
Seed dressing with Baytan and barley yield in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Kingwell R (1994) Effects of tactical responses and risk aversion on farm wheat supply. Review of Marketing and Agricultural Economics 62, 29–42.

Kingwell R (1997) The role of tactical decision-making and risk aversion in affecting a farmer’s response to price stabilization. In ‘Proceedings AARES 1997 Conference (41st)’. 22–24 January, Gold Coast, Qld. (Australian Agricultural and Resource Economics Society) Available at: https://ageconsearch.umn.edu/search?cc=408&ln=en&jrec=41

Kingwell R, Fuchsbichler A (2011) The whole-farm benefits of controlled traffic farming: an Australian appraisal. Agricultural Systems 104, 513–521.
The whole-farm benefits of controlled traffic farming: an Australian appraisal.Crossref | GoogleScholarGoogle Scholar |

Kingwell R, Pannell D (2005) Economic trends and drivers affecting the wheatbelt of Western Australia to 2030. Australian Journal of Agricultural Research 56, 553–561.
Economic trends and drivers affecting the wheatbelt of Western Australia to 2030.Crossref | GoogleScholarGoogle Scholar |

Kingwell RS, Pannell D, Robinson S (1993) Tactical responses to seasonal conditions in whole-farm planning in Western Australia. Agricultural Economics 8, 211–226. http://purl.umn.edu/173075

Kirkegaard JA, Hunt JR, McBeath TM, Lilley JM, Moore A, Verburg K, Robertson M, Oliver Y, Ward PR, Milroy S, Whitbread AM (2014) Improving water productivity in the Australian grains industry—a nationally coordinated approach. Crop & Pasture Science 65, 583–601.
Improving water productivity in the Australian grains industry—a nationally coordinated approach.Crossref | GoogleScholarGoogle Scholar |

Kirkegaard JA, Lille JM, Morrison MJ (2016) Drivers of trends in Australian canola productivity and future prospects. Crop & Pasture Science 67, i–ix.
Drivers of trends in Australian canola productivity and future prospects.Crossref | GoogleScholarGoogle Scholar |

Kwak YS, Weller DM (2013) Take-all of wheat and natural disease suppression. A review. Plant Pathology Journal 29, 125–135.
Take-all of wheat and natural disease suppression. A review.Crossref | GoogleScholarGoogle Scholar | 25288939PubMed |

Ladha JK, Pathak H, Krupnik TJ, Six J, van Kessel C (2005) Efficiency of fertiliser nitrogen in cereal production: retrospects and prospects. Advances in Agronomy 87, 85–156.
Efficiency of fertiliser nitrogen in cereal production: retrospects and prospects.Crossref | GoogleScholarGoogle Scholar |

Lawes RA (2015) Crop sequences in modern Australian farming systems. Crop & Pasture Science 66, i–ii.
Crop sequences in modern Australian farming systems.Crossref | GoogleScholarGoogle Scholar |

Lawes R, Renton M (2015) Gaining insight into the risks, returns and value of perfect knowledge for crop sequences by comparing optimal sequences with those proposed by agronomists. Crop & Pasture Science 66, 622–633.
Gaining insight into the risks, returns and value of perfect knowledge for crop sequences by comparing optimal sequences with those proposed by agronomists.Crossref | GoogleScholarGoogle Scholar |

Lawes RA, Gupta VVSR, Kirkegaard JA, Roget DK (2013) Evaluating the contribution of take-all control to the break-crop effect in wheat. Crop & Pasture Science 64, 563–572.
Evaluating the contribution of take-all control to the break-crop effect in wheat.Crossref | GoogleScholarGoogle Scholar |

Lawrence L (2009) Protecting germinating grain crops in southern Australia. Outlooks on Pest Management 20, 178–181.
Protecting germinating grain crops in southern Australia.Crossref | GoogleScholarGoogle Scholar |

Le Gall M, Tooker JF (2017) Developing ecologically based pest management programs for terrestrial molluscs in field and forage crops. Journal of Pest Science 90, 825.
Developing ecologically based pest management programs for terrestrial molluscs in field and forage crops.Crossref | GoogleScholarGoogle Scholar |

Leach JE, Stevenson HJ, Rainbow AJ, Mullen LA (1999) Effects of high plant populations on the growth and yield of winter oilseed rape (Brassica napus). The Journal of Agricultural Science 132, 173–180.
Effects of high plant populations on the growth and yield of winter oilseed rape (Brassica napus).Crossref | GoogleScholarGoogle Scholar |

Leather SR, Atanasova D (2017) Without up-to-date pest thresholds sustainable agriculture is nothing but a pipe-dream. Agricultural and Forest Entomology 19, 341–343.
Without up-to-date pest thresholds sustainable agriculture is nothing but a pipe-dream.Crossref | GoogleScholarGoogle Scholar |

Legizamon E, Colombo ME, Salinas A, Severin C (1980) Models of emergence of nineteen weed species. Malezas 8, 3–11.

Leonard  L (1993 ) Managing for stubble retention. Bulletin 4271. Department of Agriculture Western Australia, Perth, W. Aust. Available at: https://researchlibrary.agric.wa.gov.au/cgi/viewcontent.cgi?article=1137&context=bulletins

Li Y, Chen D, Walker CN, Angus JF (2010) Estimating the nitrogen status of crops using a digital camera. Field Crops Research 118, 221–227.
Estimating the nitrogen status of crops using a digital camera.Crossref | GoogleScholarGoogle Scholar |

Limpert E, Godet F, Müller K (1999) Dispersal of cereal mildew across Europe. Agricultural and Forest Meteorology 97, 293–308.
Dispersal of cereal mildew across Europe.Crossref | GoogleScholarGoogle Scholar |

Littler JW (1963) Effect of time and rate of application of urea to wheat. Queensland Journal of Agriculture and Animal Science 20, 323–344.

Liu A, Ridsdill-Smith TJ (2001) Comparisons of feeding damage by redlegged earth mite Halotydeus destructor (Tucker) (Acari: Penthaleidae) to different grain legume species as an indicator of potentially resistant lines. In ‘Acarology: Proceedings 10th International Congress’. (Eds RB Halliday, DE Walter, HC Proctor, RA Norton, MJ Colloff) pp. 83–99. (CSIRO Publishing: Melbourne)

Llewellyn RS, D’Emden F (2009) Adoption of no-till cropping practices in Australian grain growing regions. Report for SA No-till Farmers Association and CAAANZ. Grains Research and Development Corporation, Canberra, ACT. Available at: https://grdc.com.au/__data/assets/pdf_file/0016/125206/grdcadoptionofnotillcroppingpracticesreportpdf.pdf

Loss S, Haddad A, Khalil Y, Alrijabo AS, Feindel D, Piggin C (2015) Evolution and adoption of conservation agriculture in the Middle East. In ‘Conservation agriculture’. (Eds M Farooq, KHM Siddique) pp. 197–224. (Springer International Publishing: Cham, Switzerland)

Loughman R, Wright D, MacLeod W, Bhathal J, Thackray D (2000) Diseases. In ‘The wheat book–principles and practice’. Bulletin 4443. (Eds WK Anderson, JR Garlinge) pp. 201–228. (Agriculture Western Australia: Perth, W. Aust.)

Loughman R, Jayasena KW, Majewski J (2005) Yield loss and fungicide control of stem rust of wheat. Australian Journal of Agricultural Research 56, 91–96.
Yield loss and fungicide control of stem rust of wheat.Crossref | GoogleScholarGoogle Scholar |

Ludwig F, Milroy SP, Asseng S (2009) Impacts of recent climate change on wheat production systems in Western Australia. Climatic Change 92, 495–517.
Impacts of recent climate change on wheat production systems in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Luo Q, Kathuria A (2013) Modelling the response of wheat grain yield to climate change: a sensitivity analysis. Theoretical and Applied Climatology 111, 173–182.
Modelling the response of wheat grain yield to climate change: a sensitivity analysis.Crossref | GoogleScholarGoogle Scholar |

Macfadyn S, Hill M (2017) Arthropod pests of Australian canola during crop emergence: IPM and future directions. In ‘Integrated management of insect pests on canola and other brassica oilseed crops’. (Ed. GVP Reddy) pp. 350–372. (CABI Publishing: Wallingford, UK)

MacLeod WJ, MacNish GC, Horn CW (1993) Manipulation of ley pastures with herbicides to control take-all. Australian Journal of Agricultural Research 44, 1235–1244.
Manipulation of ley pastures with herbicides to control take-all.Crossref | GoogleScholarGoogle Scholar |

Madin R (1993) Non-chemical weed control. In ‘Management of agricultural weeds in Western Australia’. Bulletin 4243. (Eds J Dodd, RJ Martin, KM Howes) pp. 47–49. (Western Australian Department of Agriculture: Perth, W. Aust.)

Marroni MV, Viljanen-Rollinson SLH, Butler RC, Deng Y (2006) Fungicide timing for the control of septoria tritici blotch of wheat. New Zealand Plant Protection 59, 160–165.
Fungicide timing for the control of septoria tritici blotch of wheat.Crossref | GoogleScholarGoogle Scholar |

Mason MG (1968) Sources of nitrogen for cereals—their characteristics and costs. Journal of Agriculture of Western Australia 9, 556–558.

Mason MG (1977) Sources of nitrogen for cereals—urea, ammonium nitrate or sulphate of ammonia? Journal of Agriculture of Western Australia 18, 15–19.

Mathre DE (1997) ‘Compendium of barley diseases.’ (The American Phytopathological Society: St Paul, MN, USA)

McDonald BA, Mundt CC (2016) How knowledge of pathogen population biology informs management of septoria tritici blotch. Phytopathology 106, 948–955.
How knowledge of pathogen population biology informs management of septoria tritici blotch.Crossref | GoogleScholarGoogle Scholar | 27111799PubMed |

Michael P (2002) Cereal aphids and direct feeding damage to cereals. In ‘2002 cereals update’. (Ed. R Jettner) pp. 55–56. (Department of Agriculture: Perth, W, Aust.) https://researchlibrary.agric.wa.gov.au/cgi/viewcontent.cgi?article=1014&context=crop_up (accessed 10 June 2020)

Micic S (2005) Identification and cultural control of insect and allied pests of canola. Bulletin 4650. Department of Agriculture Western Australia, Perth, W. Aust.

Micic S, Michael P (2005) Insect contamination of cereal grain at harvest. In ‘Proceedings 2005 Agribusiness Crop Updates: cereal updates’. (Ed. R Horne) pp. 81–82. (Department of Agriculture Western Australia: Perth, W. Aust.)

Micic S, Matson P, Dore T (2006) How to avoid insect contamination in cereal grain at harvest. In ‘Proceedings 2006 Agribusiness Crop Updates: cereal updates’. (Ed. S Penny) pp. 91–92. (Department of Agriculture Western Australia: Perth, W, Aust.)

Micic S, Dore A, Strickland G (2007) The effect of rotation crops, trash retention and prophylactic sprays on arthropod abundance in a following canola crop. In ‘Proceedings 2007 Agribusiness Crop Updates: lupins, pulses and oilseeds updates’. (Ed. W Parker) pp. 75–78. (Department of Agriculture and Food: Perth, W. Aust.)

Micic S, Hoffmann AA, Strickland G, Weeks AR, Bellati J, Henry K, Nash MA, Umina PA (2008) Pests of germinating grain crops in southern Australia: an overview of their biology and management options. Australian Journal of Experimental Agriculture 48, 1560–1573.
Pests of germinating grain crops in southern Australia: an overview of their biology and management options.Crossref | GoogleScholarGoogle Scholar |

Miller RH, Pike KS (2002) Insects in wheat-based systems. In ‘Bread wheat: improvement and production, plant production and protection series. No. 30’. (Eds BC Curtis, S Rajaram S, H Gómez Macpherson) pp. 367–393. (Food and Agriculture Organization of the United Nations: Rome)

Minkey DM, Moore JH (1996) Estimating dose response curves for predicting glyphosate use in Australia. In ‘Proceedings Eleventh Australian Weeds Conference’. (Ed. RCH Shepherd) pp. 166–169. (Weed Science Society of Victoria: Melbourne) Available at: http://caws.org.nz/old-site/awc/1996/awc199611661.pdf

Monjardino M, Pannell DJ, Powles SB (2003) Multispecies resistance and integrated management: a bioeconomic model for integrated management of rigid ryegrass (Lolium rigidum) and wild radish (Raphanus raphanistrum). Weed Science 51, 798–809.
Multispecies resistance and integrated management: a bioeconomic model for integrated management of rigid ryegrass (Lolium rigidum) and wild radish (Raphanus raphanistrum).Crossref | GoogleScholarGoogle Scholar |

Moore JH (1979) Influence of weed species and density on the yield of crops. In ‘Proceedings Western Australian Weeds Conference’. (Ed. R Madin) pp. 92–94. (The Weed Society of Western Australia: Perth, W. Aust.) Available at: www.herbiguide.com.au/Downloads/InfluenceOfWeedSpeciesAndDensityOnTheYieldOfCrops.pdf

Moore JH, Minkey DM (1997) HerbiRate: An excel spreadsheet for adjusting rates based on growing conditions. HerbiGuide, Albany, W. Aust. Available at: http://www.herbiguide.com.au/Downloads/HerbiRate.xls

Moore CB, Moore JH (2020) HerbiGuide—the pesticide expert on a disk. HerbiGuide, Albany, W. Aust. Available at: http://www.herbiguide.com.au/

Moore JH, Wheeler J (2020) Southern weeds and their control. 4th edn. Bulletin 4914. Department of Primary Industries and Regional Development, Perth, W. Aust.

Moore J, Babativa Rodriguez C, Micic S (2019) Grain Cam: a device for detecting contaminants of grain as it is being harvested. GRDC Update Papers 26 Feb. 2019. Grains Research and Development Corporation, Canberra, ACT. Available at: Available at: https://grdc.com.au/resources-and-publications/grdc-update-papers/tab-content/grdc-update-papers/2019/02/graincam-a-device-for-detecting-contaminants-in-grain-as-it-is-being-harvested

Muñoz-Huerta RF, Guevara-Gonzalez RG, Contreras-Medina LM, Torres-Pacheco I, Prado-Olivarez J, Ocampo-Velazquez RV (2013) A review of methods for sensing the nitrogen status in plants: advantages, disadvantages and recent advances. Sensors 13, 10823–10843.
A review of methods for sensing the nitrogen status in plants: advantages, disadvantages and recent advances.Crossref | GoogleScholarGoogle Scholar | 23959242PubMed |

Murray TD, Parry DW, Cattlin ND (2009) ‘Diseases of small grain cereal crops. A colour handbook.’ (Manson Publishing: London)

Murray D, Clark M, Ronning D (2012) ‘Current and potential costs of invertebrate pests in grain crops.’ Publication AEP001. (Grains Research and Development Corporation: Canberra, ACT)

Nichols PGH, Dear BS, Hackney BF, Craig AD, Evans PM, de Koning CT, Foster KJ, Barbetti MJ, You MP, Micic S (2009) New subterranean clovers with reduced cotyledon susceptibility to red-legged earth mites. SABRAO Journal of Breeding and Genetics 41, 16.

Nichols PGH, Revell CK, Humphries AW, Howie JH, Hall EJ, Sandral GA, Ghamkhar K, Harris CA (2012) Temperate pasture legumes in Australia—their history, current use and future prospects. Crop & Pasture Science 63, 691–725.
Temperate pasture legumes in Australia—their history, current use and future prospects.Crossref | GoogleScholarGoogle Scholar |

Nix HA, Fitzpatrick EA (1969) An index of crop water stress related to wheat and grain sorghum yields. Agricultural Meteorology 6, 321–337.
An index of crop water stress related to wheat and grain sorghum yields.Crossref | GoogleScholarGoogle Scholar |

Nordblom T (2003) Putting biological reality into economic assessments of biocontrol. In ‘Improving the selection, testing and evaluation of weed biocontrol agents. Proceedings Biological Control of Weeds Symposium and Workshop’. Series No. 7. (Eds H Spafford-Jacob, D Briese) pp. 75–85. (CRC for Australian Weed Management: Perth, W. Aust.)

Nordblom TL, Ahmed AH, Miller SF, Glenn DM (1985) Long-run evaluation of fertilization strategies for dryland wheat in northcentral Oregon: simulation analysis. Agricultural Systems 18, 133–153.
Long-run evaluation of fertilization strategies for dryland wheat in northcentral Oregon: simulation analysis.Crossref | GoogleScholarGoogle Scholar |

Nyborg M, Hennig AMF (1969) Field experiments with different placements of fertilisers for barley, flax and rapeseed. Canadian Journal of Soil Science 49, 79–88.
Field experiments with different placements of fertilisers for barley, flax and rapeseed.Crossref | GoogleScholarGoogle Scholar |

O’Connell M, Pannell DJ, French RJ (2003) Are high lupin seeding rates more risky in Western Australian wheatbelt? Australian Journal of Experimental Agriculture 43, 1137–1142.
Are high lupin seeding rates more risky in Western Australian wheatbelt?Crossref | GoogleScholarGoogle Scholar |

Ophel-Keller K, McKay A, Hartley D (2008) Development of a routine DNA-based testing service for soilborne diseases in Australia. Australasian Plant Pathology 37, 243–253.
Development of a routine DNA-based testing service for soilborne diseases in Australia.Crossref | GoogleScholarGoogle Scholar |

Owen M, Powles SB (2018) Current levels of herbicide resistance in key weed species in the WA grain belt. GRDC Update Papers 26 Feb. 2018. Grains Research and Development Corporation, Canberra, ACT. Available at: https://grdc.com.au/resources-and-publications/grdc-update-papers/tab-content/grdc-update-papers/2018/02/current-levels-of-herbicide-resistance-in-key-weed-species

Owen MJ, Martinez NJ, Powles SB (2014) Multiple herbicide‐resistant Lolium rigidum (annual ryegrass) now dominates across the Western Australian grain belt. Weed Research 54, 314–324.
Multiple herbicide‐resistant Lolium rigidum (annual ryegrass) now dominates across the Western Australian grain belt.Crossref | GoogleScholarGoogle Scholar |

Ozanne PG, Petch A (1978) The application of nutrients by foliar sprays to increase seed yields. In ‘Proceedings 8th International Colloquium on Plant Analysis and Fertiliser Problems’. (Eds AR Ferguson, RL Bieleski, IB Ferguson) pp. 361–366. (Department of Scientific and Industrial Research: Auckland, NZ)

Palta JA, Peltzer S (2001) Annual ryegrass (Lolium rigidum) reduces the uptake and utilisation of fertiliser nitrogen by wheat. Australian Journal of Agricultural Research 52, 573–581.
Annual ryegrass (Lolium rigidum) reduces the uptake and utilisation of fertiliser nitrogen by wheat.Crossref | GoogleScholarGoogle Scholar |

Park RF (2008) Breeding cereals for rust resistance in Australia. Plant Pathology 57, 591–602.
Breeding cereals for rust resistance in Australia.Crossref | GoogleScholarGoogle Scholar |

Paynter B, Trainor G, Curry J (2019) Barley variety sowing guide for Western Australia. Bulletin 4886. Department of Primary Industries and Regional Development, Perth, W. Aust. Available at: https://www.agric.wa.gov.au/barley/2018-barley-variety-sowing-guide-western-australia

Peel MC, Finlayson BL, McMahon TA (2007) Updated world map of the Koppen–Geiger climate classification. Hydrology and Earth System Sciences 11, 1633–1644.
Updated world map of the Koppen–Geiger climate classification.Crossref | GoogleScholarGoogle Scholar |

Peirce CAE, McBeath TM, Priest C, McLaughlin MJ (2019) The timing of application and inclusion of a surfactant are important for absorption and translocation of foliar phosphoric acid by wheat leaves. Frontiers in Plant Science 10, 1532.
The timing of application and inclusion of a surfactant are important for absorption and translocation of foliar phosphoric acid by wheat leaves.Crossref | GoogleScholarGoogle Scholar | 31824546PubMed |

Peverill KI, Sparrow LA, Reuter DJ (1999) ‘Soil analysis: an interpretation manual.’ (CSIRO Publishing: Melbourne)

Piening LJ (1967) Note on the effect of N, P, and K in several Alberta soils on the resistance of gateway barley to the net blotch disease. Canadian Journal of Plant Science 47, 713–715.
Note on the effect of N, P, and K in several Alberta soils on the resistance of gateway barley to the net blotch disease.Crossref | GoogleScholarGoogle Scholar |

Piggin C, Haddad A, Khalil YY, Loss S, Pala M (2015) Effects of tillage and time of sowing on bread wheat, chickpea, barley and lentil grown in rotation in rainfed systems in Syria. Field Crops Research 173, 57–67.
Effects of tillage and time of sowing on bread wheat, chickpea, barley and lentil grown in rotation in rainfed systems in Syria.Crossref | GoogleScholarGoogle Scholar |

Platz GJ, Meldrum SI, Webb NA (1999) Chemical control of seed borne diseases of barley. In ‘Proceeding 9th Australian Barley Technical Symposium’. (The Regional Institute: Gosford, NSW) Available at: http://www.regional.org.au/au/abts/1999/platzhtm.htm#TopOfPage

Poole ML, Turner NC, Young YM (2002) Sustainable cropping systems for high rainfall areas of south Western Australia. Agricultural Water Management 53, 201–211.
Sustainable cropping systems for high rainfall areas of south Western Australia.Crossref | GoogleScholarGoogle Scholar |

Radford BJ, Wilson BJ, Cartledge O, Watkins FB (1980) Effect of wheat seeding rate on wild oat competition. Australian Journal of Experimental Agriculture and Animal Husbandry 20, 77–81.
Effect of wheat seeding rate on wild oat competition.Crossref | GoogleScholarGoogle Scholar |

Ramsden MW, Kendall SL, Ellis SA, Berry PM (2017) A review of economic thresholds for invertebrate pests in UK arable crops. Crop Protection 96, 30–43.
A review of economic thresholds for invertebrate pests in UK arable crops.Crossref | GoogleScholarGoogle Scholar |

Reeves JT (1954) Some effects of spraying wheat with urea. Journal of the Australian Institute of Agricultural Science 20, 41–45.

Renton M, Lawes R, Metcalf T, Robertson M (2015) Considering long-term ecological effects on future land-use options when making tactical break-crop decisions in cropping systems. Crop & Pasture Science 66, 610–621.
Considering long-term ecological effects on future land-use options when making tactical break-crop decisions in cropping systems.Crossref | GoogleScholarGoogle Scholar |

Reuter DJ, Robinson JB (1997) ‘Plant analysis: an interpretation manual.’ 2nd edn. (CSIRO Publishing: Melbourne)

Richards RA, Hunt JR, Kirkegaard JA, Passioura JB (2014) Yield improvement and adaptation of wheat to water-limited environments in Australia—a case study. Crop & Pasture Science 65, 676–689.
Yield improvement and adaptation of wheat to water-limited environments in Australia—a case study.Crossref | GoogleScholarGoogle Scholar |

Ridsdill-Smith J, Hoffmann AA, Mangano P, Gower JM, Pavri C, Umina PA (2008) Review of strategies for control of the redlegged earth mite in Australia. Australian Journal of Experimental Agriculture 48, 1506–1513.
Review of strategies for control of the redlegged earth mite in Australia.Crossref | GoogleScholarGoogle Scholar |

Robertson MJ, Llewellyn RS, Mandel R, Lawes R, Bramley RGV, Swift L, Metz N, O’Callaghan C (2012) Adoption of variable rate fertiliser application in the Australian grains industry: status, issues and prospects. Precision Agriculture 13, 181–199.
Adoption of variable rate fertiliser application in the Australian grains industry: status, issues and prospects.Crossref | GoogleScholarGoogle Scholar |

Roques SE, Berry PM (2016) The yield response of oilseed rape to plant population density. The Journal of Agricultural Science 154, 305–320.
The yield response of oilseed rape to plant population density.Crossref | GoogleScholarGoogle Scholar |

Sadras VO, Angus JF (2006) Benchmarking water use efficiency of rainfed wheat crops in dry mega-environments. Australian Journal of Agricultural Research 57, 847–856.
Benchmarking water use efficiency of rainfed wheat crops in dry mega-environments.Crossref | GoogleScholarGoogle Scholar |

Salam MU, Salam KP, Thomas GJ, Beard C, William JM, Diggle AJ, Bowran DG (2013) Big achievement from a handful of high performing varieties: a concept in the context of crop health for wheat in Western Australia. In ‘Proceedings 19th Biennial Australasian Plant Pathology Society Conference’. Auckland, New Zealand. (Australasian Plant Pathology Society) Available at: https://www.appsnet.org/publications/proceedings/APPS%202013%20Handbook.pdf

Schiere JB, Baumhardt RL, Van Keulen H, Whitbread AM, Bruinsma AS, Goodchild AV, Gregorini P, Slingerland MA, Weidemann-Hartwell B (2006) Mixed crop–livestock systems in semi-arid regions. In ‘Dryland agriculture’. Agronomy Monograph No. 23. (Eds GA Peterson, PW Unger, WA Payne) pp. 227–292. (ASA, CSSA, SSSA: Madison, WI, USA)

Schmidt CP, Belford RK (1993) A comparison of different tillage-seeding systems: the interaction of tillage and time of sowing on sandplain soils in Western Australia. Australian Journal of Experimental Agriculture 33, 895–900.
A comparison of different tillage-seeding systems: the interaction of tillage and time of sowing on sandplain soils in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Scott BJ, Eberbach PL, Evans J, Wade LJ (2010) ‘Stubble retention in cropping systems in southern Australia: benefits and challenges.’ Monograph No. 1. (Graham Centre for Agricultural Innovation, EH Graham Centre: Wagga Wagga, NSW) Available at: https://cdn.csu.edu.au/__data/assets/pdf_file/0007/922723/stubble-retention.pdf

Serraj R, Siddique KHM (2012) Conservation agriculture in dry areas. Field Crops Research 132, 1–6.
Conservation agriculture in dry areas.Crossref | GoogleScholarGoogle Scholar |

Seymour M (2011) Defining economic plant densities of open pollinated and hybrid canola in WA. In ‘Proceedings 2011 Agribusiness Crop Updates’. (Eds J Paterson, C Nicolls) pp. 213–216. (Department of Agriculture and Food Western Australia: Perth, W. Aust.)

Seymour M, Brennan RF (1995) Nutrient sprays applied to the foliage of narrow-leafed lupins (Lupinus angustifolius L.) during flowering and podding do not increase seed yield. Australian Journal of Experimental Agriculture 35, 381–383.
Nutrient sprays applied to the foliage of narrow-leafed lupins (Lupinus angustifolius L.) during flowering and podding do not increase seed yield.Crossref | GoogleScholarGoogle Scholar |

Seymour M, Kirkegaard JA, Peoples MB, White PF, French RJ (2012) Break-crop benefits to wheat in Western Australia—insights from over three decades of research. Crop & Pasture Science 63, 1–16.
Break-crop benefits to wheat in Western Australia—insights from over three decades of research.Crossref | GoogleScholarGoogle Scholar |

Seymour M, Sprigg S, French B, Bucat J, Malik R, Harries M (2016) Nitrogen responses of canola in low and medium rainfall zones of Western Australia. Crop & Pasture Science 67, 450–466.
Nitrogen responses of canola in low and medium rainfall zones of Western Australia.Crossref | GoogleScholarGoogle Scholar |

Shackley  BPaynter  BTroup  GBucat  JSeymour  MBlake  A (2020 ) 2020 Western Australian crop sowing guide. Bulletin 4910. Department of Primary Industries and Regional Development, Perth, W. Aust. https://www.agric.wa.gov.au/barley/2020-western-australian-crop-sowing-guide

Shankar M, Mather D, Jorgensen D, Golzar H, Chalmers K, Hollaway G, McLean M, Neate S, Loughman R (2015) Germplasm enhancement for resistance to Pyrenophora tritici-repentis in wheat. In ‘Advances in wheat genetics: from genome to field’. (Eds Y Ogihara, S Takumi, H Handa) pp. 193–199. (Springer: Tokyo)

Sharma D, Anderson W (2004) Small grain screenings in wheat: interactions of cultivars, with season, site and management practices. Australian Journal of Agricultural Research 55, 797–809.
Small grain screenings in wheat: interactions of cultivars, with season, site and management practices.Crossref | GoogleScholarGoogle Scholar |

Sharma D, D’Antuono MF, Anderson WK, Shackley BJ, Zaicou-Kunesch CM, Amjad M (2008) Variability of optimum sowing time for wheat yield in Western Australia. Australian Journal of Agricultural Research 59, 958–970.
Variability of optimum sowing time for wheat yield in Western Australia.Crossref | GoogleScholarGoogle Scholar |

Shipton WA (1966) Effect of net blotch infection on grain yield and quality. Australian Journal of Experimental Agriculture and Animal Husbandry 6, 437–440.
Effect of net blotch infection on grain yield and quality.Crossref | GoogleScholarGoogle Scholar |

Shipton WA, Kahn TN, Boyd WJR (1973) Net blotch of barley. Reviews in Plant Pathology 52, 269–290.

Simkin MB, Wheeler BEJ (1974) The development of Puccinia hordei on barley cv. Zephyr. Annals of Applied Biology 78, 225–235.
The development of Puccinia hordei on barley cv. Zephyr.Crossref | GoogleScholarGoogle Scholar |

Simpson NL, Brennan RF, Anderson WK (2016) Grain yield increases in wheat and barley to nitrogen applied after transient waterlogging in the high rainfall cropping zone of Western Australia. Journal of Plant Nutrition 39, 974–992.
Grain yield increases in wheat and barley to nitrogen applied after transient waterlogging in the high rainfall cropping zone of Western Australia.Crossref | GoogleScholarGoogle Scholar |

Skoropad WP (1960) Barley scald in the prairie provinces of Canada. Commonwealth Phytopathological News 6, 25–27.

Smith ST, Toms WJ (1958) Manganese deficiency in the cereal growing areas of Western Australia. Journal of the Department of Agriculture Western Australia (3rd series) 7, 65–70.

Solomon PS, Lowe RGT, Tan KC, Waters ODC, Oliver RP (2006) Stagonospora nodorum: cause of stagonospora nodorum blotch of wheat. Molecular Plant Pathology 7, 147–156.
Stagonospora nodorum: cause of stagonospora nodorum blotch of wheat.Crossref | GoogleScholarGoogle Scholar | 20507435PubMed |

Somerville GJ, Powles SB, Walsh MJ, Renton M (2018) Modelling the impact of harvest weed seed control on herbicide resistance evolution. Weed Science 66, 395–403.
Modelling the impact of harvest weed seed control on herbicide resistance evolution.Crossref | GoogleScholarGoogle Scholar |

Sommer R, Piggin C, Feindel D, Ansar M, van Delden L, Shimonaka K, Abdalla J, Douba O, Estefan G, Haddad A, Haj-Abdo R, Hajdibo A, Hayek P, Khalil Y, Khoder A, Ryan J (2014) Effects of zero tillage and residue retention on soil quality in the Mediterranean region of northern Syria. Open Journal of Soil Science 4, 109–125.
Effects of zero tillage and residue retention on soil quality in the Mediterranean region of northern Syria.Crossref | GoogleScholarGoogle Scholar |

Stephens DJ, Lyons TJ (1998) Variability and trends in sowing dates across the Australian wheatbelt. Australian Journal of Agricultural Research 49, 1111–1118.
Variability and trends in sowing dates across the Australian wheatbelt.Crossref | GoogleScholarGoogle Scholar |

Stephens DJ, Lyons TJ, Lamond MH (1989) A simple model to forecast wheat yield in Western Australia. Journal of the Royal Society of Western Australia 71, 77–81.

Stern VM (1973) Economic thresholds. Annual Review of Entomology 18, 259–280.
Economic thresholds.Crossref | GoogleScholarGoogle Scholar |

Te Beest DE, Paveley ND, Shaw MW, van den Bosch F (2008) Disease–weather relationships of powdery mildew and yellow rust on winter wheat. Phytopathology 98, 609–617.
Disease–weather relationships of powdery mildew and yellow rust on winter wheat.Crossref | GoogleScholarGoogle Scholar | 18943230PubMed |

Teng PS, Close RC (1978) Effect of temperature and uredinium density on urediniospores production, latent period, and infectious period of Puccinia hordei Otth. New Zealand Journal of Agricultural Research 21, 287–296.
Effect of temperature and uredinium density on urediniospores production, latent period, and infectious period of Puccinia hordei Otth.Crossref | GoogleScholarGoogle Scholar |

Thackray DJ, Diggle AJ, Jones RAC (2004) Forecasting aphid outbreaks and epidemics of cucumber mosaic virus in narrow-leafed lupins (Lupinus angustifolius). Virus Research 100, 67–82.
Forecasting aphid outbreaks and epidemics of cucumber mosaic virus in narrow-leafed lupins (Lupinus angustifolius).Crossref | GoogleScholarGoogle Scholar | 15036837PubMed |

Thomas G, Beard C, Smith A, Jayasena K, Kelly S (2008) Leaf disease management in continuous barley in the northern and central grainbelt of WA. 2008 Cereal Updates. Department of Agriculture, Perth, W. Aust. pp. 68–72. Available at: http://citeseerx.ist.psu.edu/viewdoc/download;jsessionid=082D80E126A033E8A93A609F53CE05CD?doi=10.1.1.626.2626&rep=rep1&type=pdf (accessed 10 June 2020).

Thomas G, Beard C, Jayasena K, Hills A, Bradley J, Smith A (2017) Fungicides at seeding for management of cereal foliar diseases: powdery mildew in wheat. GRDC Update Papers 27 February 2017. Grains Research and Development Corporation, Canberra, ACT. Available at: https://grdc.com.au/resources-and-publications/grdc-update-papers/tab-content/grdc-update-papers/2017/02/fungicides-at-seeding-for-management-of-cereal-foliar-diseases-powdery-mildew-in-wheat

Tokatlidis IS (2014) Addressing the yield by density interaction is a prerequisite to bridge the yield gap of rainfed wheat. Annals of Applied Biology 165, 27–42.

Tosti G, Farneselli M, Benincasa P, Guidicci M (2016) Nitrogen fertilization strategies for organic wheat production: crop yield and nitrate leaching. Agronomy Journal 108, 770–781.
Nitrogen fertilization strategies for organic wheat production: crop yield and nitrate leaching.Crossref | GoogleScholarGoogle Scholar |

Tucker MA, Lopez-Ruiz F, Jayasena KW, Oliver RP (2015) Origin of fungicide resistant barley powdery mildew in West Australia—lessons to be learned. In ‘Fungicide resistance in plant pathogens: principles and a guide to practical management’. (Eds H Ishii, DW Holloman) pp. 329–340. (Springer: Tokyo)

Turner NC (2004) Agronomic options for improving rainfall use efficiency of crops in dryland farming systems. Water-Saving Agriculture Special Edition. Journal of Experimental Botany 55, 2413–2435.

Umina PA, Micic S, Fagan L (2011) Control of insect and mite pests in grains—insecticide resistance and integrated pest management (IPM). In ‘Proceedings 2011 Agribusiness Crop Updates’. (Eds J Paterson, C Nicholls) pp. 98–100. (Department of Agriculture Western Australia: Perth, W. Aust.)

van den Berg CGJ, Rossnagel BG (1990) Effect of temperature and leaf wetness period on conidium germination and infection of barley by Pyrenophora teres. Canadian Journal of Plant Pathology 12, 263–266.
Effect of temperature and leaf wetness period on conidium germination and infection of barley by Pyrenophora teres.Crossref | GoogleScholarGoogle Scholar |

van Heerwaarden AF, Angus JF, Richards RA, Farquhar GD (1998) “Haying off’, the negative grain yield response of dryland wheat to nitrogen fertilizer. II. Carbohydrate and protein dynamics. Australian Journal of Agricultural Research 49, 1083–1093.

Vigneau N, Ecarnot M, Rabatel G, Roumet P (2011) Potential of field hyperspectral imaging as a non-destructive method to assess leaf nitrogen content in wheat. Field Crops Research 122, 25–31.
Potential of field hyperspectral imaging as a non-destructive method to assess leaf nitrogen content in wheat.Crossref | GoogleScholarGoogle Scholar |

Virgona JM, Gummer F, Angus JF (2006) Effects of grazing on wheat growth, yield, development, water use and nitrogen use. Australian Journal of Agricultural Research 57, 1307–1319.
Effects of grazing on wheat growth, yield, development, water use and nitrogen use.Crossref | GoogleScholarGoogle Scholar |

Wallace MM (1970) The influence of temperature on post-diapause development and survival of aestivating eggs of Halotydeus destructor (Acari: Eupodidae). Australian Journal of Zoology 18, 315–329.
The influence of temperature on post-diapause development and survival of aestivating eggs of Halotydeus destructor (Acari: Eupodidae).Crossref | GoogleScholarGoogle Scholar |

Wallwork H, Preece P, Cotterill PJ (1992) Puccinia hordei on barley and Ornithogalum umbellatum in South Australia. Australasian Plant Pathology 21, 95–97.
Puccinia hordei on barley and Ornithogalum umbellatum in South Australia.Crossref | GoogleScholarGoogle Scholar |

Walsh M (2016) Weed management in dryland cropping systems. In ‘Innovations in dryland agriculture’. (Eds M Farooq, KHM Siddique) pp. 99–114. (Springer International Publishing: Cham, Switzerland)

Walsh M, Minkey DM (2006) Wild radish (Raphanus raphanistrum L.) development and seed production in response to time of emergence, crop topping and sowing rate of wheat. Plant Protection Quarterly 21, 25–29.

Walsh MJ, Devlin RD, Powles SB (2004) Potential for preseason herbicide application to prevent weed emergence in the subsequent growing season. 1. Identification and evaluation of possible herbicides. Weed Technology 18, 228–235.
Potential for preseason herbicide application to prevent weed emergence in the subsequent growing season. 1. Identification and evaluation of possible herbicides.Crossref | GoogleScholarGoogle Scholar |

Walsh M, Ouzman J, Newman P, Powles S, Llewellyn R (2017) High levels of adoption indicate that harvest weed seed control is now an established weed control practice in Australian cropping. Weed Technology 31, 341–347.
High levels of adoption indicate that harvest weed seed control is now an established weed control practice in Australian cropping.Crossref | GoogleScholarGoogle Scholar |

Williams S (2017) Key insect and mite pests of Australian cotton. In ‘Cotton pest management guide’. (Eds S Saas, R Redfern) pp. 5–9. (Cotton Research and Development Corporation: Narrabri, NSW)

Williams DG, Il’ichev AL (2003) Integrated pest management in Australia. In ‘Integrated pest management in the global arena’. (Eds KM Maredia, D Dakouo, D Mota-Sanchez) pp. 371–384. (CABI Publishing: Wallingford, UK)

Wilson JM (1989) Leaf diseases of wheat and time of sowing. Journal of the Department of Agriculture, Western Australia: Series 4 30, art. 12.

Yenish J, Doll J, Buhler D (1992) Effects of tillage on vertical distribution and viability of weed seed in soil. Weed Science 40, 429–433.
Effects of tillage on vertical distribution and viability of weed seed in soil.Crossref | GoogleScholarGoogle Scholar |

Yigezu Y, Mugera A, El-Shater T, Piggin C, Haddad A, Khalil YY, Loss S (2014) Explaining adoption and measuring impacts of conservation agriculture on productive efficiency, income, poverty and food security in Syria. In ‘Conservation agriculture’. (Eds M Farooq, KHM Siddique) pp. 225–247. (Springer International Publishing: Cham, Switzerland)

Zadoks JC, Chang TT, Konzak CF (1974) A decimal code for the growth stages of cereals. Weed Research 14, 415–421.
A decimal code for the growth stages of cereals.Crossref | GoogleScholarGoogle Scholar |

Zaicou-Kunesch C, Trainor G, Shackley B, Curry J, Nicol D, Shankar M, Huberli D, Thomas G, Dhammu H (2018) Wheat variety sowing guide for Western Australia. Bulletin 4881. Department of Primary Industries and Regional Development, Perth. Available at: https://www.agric.wa.gov.au/grains-research-development/2018-wheat-variety-sowing-guide-western-australia