8
Crop Protection 26 (2007) 312–319 Spray formulation efficacy—holistic and futuristic perspectives Jerzy A. Zabkiewicz Plant Protection Chemistry NZ, 49 Sala Street, Rotorua, New Zealand Received 10 January 2005; accepted 17 August 2005 Abstract This overview considers the use of liquid agrichemical sprays for different applications, their complexities and current limitations, as well as future requirements needed to increase overall efficacy. This has been done by considering the characteristics and limitations in each of the steps of deposition, retention, uptake and translocation. It has identified the lack of understanding of the plant component in relation to the processes involved in retention, uptake and translocation. Satisfactory progress will not be made until appropriate models are developed for each process and integrated into a comprehensive agrichemical efficacy system. r 2006 Elsevier Ltd. All rights reserved. Keywords: Spray formulation efficacy; Deposition; Retention; Uptake; Translocation; Models 1. Introduction The challenges facing users of agrichemicals have been increasing in complexity over recent years. On the one hand consumers require the highest quality of produce, on the other regulators insist on safety (to the consumer from residues) and risk reduction (to the operator or environ- ment). Assuming dependence on agrichemicals will con- tinue, then clear and reasonable operational procedures must be developed by agrichemical producers, equipment manufacturers, applicators, and by the regulators, to respond to the new criteria. As quoted recently ‘‘we have gone from the goal of pest death, to the more elaborate terms of yield increases, quality benefits, resistance management, IPM goal, and combinational tactics, and now towards the more sustainable and bio-intensive levels to include whole farm and agro-ecosystems approaches’’ (Hall, 2004). The requirement to increase biological efficacy and reduce detrimental environmental effects can only be met by improved spray application and formulation technolo- gies. They must be considered together as they are linked inextricably: the proportion of product intercepted by the target will vary with plant growth stage and crop type. There are many reviews on the choice and selection of nozzles and spraying systems for specific applications. More importantly, there is now an awareness that these have to be considered in relation to the spray formulation, as droplet size can vary significantly with the same spray formulation, if different nozzles or operating conditions are used (Combellack, 2004). However, although some gen- eralisations can be made, we are still a long way from being able to predict appropriate combinations accurately. In particular, the characteristics, susceptibility, receptivity, or whatever term may be applied, of the target, whether plant or other organism, is still poorly understood and may still be the dominant variable in the overall efficacy processes. An illustration of possible overall spray efficiency for an herbicide application is given in Table 1. This is a situation where the compounding effects of poor to good efficiency in individual steps of the deposition:plant retention:upta- ke:translocation interactions will give a very variable system efficiency. When individual steps are optimised, the amount received by the target site can be over a hundred and 50-fold more than the worst case. The study of spray formulation efficacy has evolved from empirical tests and field trials, through laboratory screening methods, to more detailed fundamental studies for each of the specific steps. Pesticide efficacy can be defined, and the dose received by the target site or organism modelled as a function of deposition, retention, ARTICLE IN PRESS www.elsevier.com/locate/cropro 0261-2194/$ - see front matter r 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.cropro.2005.08.019 Tel.: +64 7 343 5491; fax: +64 7 343 5811. E-mail address: [email protected].

Spray formulation efficacy—holistic and futuristic perspectives

Embed Size (px)

Citation preview

ARTICLE IN PRESS

0261-2194/$ - se

doi:10.1016/j.cr

�Tel.: +64 7

E-mail addr

Crop Protection 26 (2007) 312–319

www.elsevier.com/locate/cropro

Spray formulation efficacy—holistic and futuristic perspectives

Jerzy A. Zabkiewicz�

Plant Protection Chemistry NZ, 49 Sala Street, Rotorua, New Zealand

Received 10 January 2005; accepted 17 August 2005

Abstract

This overview considers the use of liquid agrichemical sprays for different applications, their complexities and current limitations, as

well as future requirements needed to increase overall efficacy. This has been done by considering the characteristics and limitations in

each of the steps of deposition, retention, uptake and translocation. It has identified the lack of understanding of the plant component in

relation to the processes involved in retention, uptake and translocation. Satisfactory progress will not be made until appropriate models

are developed for each process and integrated into a comprehensive agrichemical efficacy system.

r 2006 Elsevier Ltd. All rights reserved.

Keywords: Spray formulation efficacy; Deposition; Retention; Uptake; Translocation; Models

1. Introduction

The challenges facing users of agrichemicals have beenincreasing in complexity over recent years. On the onehand consumers require the highest quality of produce, onthe other regulators insist on safety (to the consumer fromresidues) and risk reduction (to the operator or environ-ment). Assuming dependence on agrichemicals will con-tinue, then clear and reasonable operational proceduresmust be developed by agrichemical producers, equipmentmanufacturers, applicators, and by the regulators, torespond to the new criteria. As quoted recently ‘‘we havegone from the goal of pest death, to the more elaborateterms of yield increases, quality benefits, resistancemanagement, IPM goal, and combinational tactics, andnow towards the more sustainable and bio-intensive levelsto include whole farm and agro-ecosystems approaches’’(Hall, 2004).

The requirement to increase biological efficacy andreduce detrimental environmental effects can only be metby improved spray application and formulation technolo-gies. They must be considered together as they are linkedinextricably: the proportion of product intercepted by thetarget will vary with plant growth stage and crop type.

e front matter r 2006 Elsevier Ltd. All rights reserved.

opro.2005.08.019

343 5491; fax: +64 7 343 5811.

ess: [email protected].

There are many reviews on the choice and selection ofnozzles and spraying systems for specific applications.More importantly, there is now an awareness that thesehave to be considered in relation to the spray formulation,as droplet size can vary significantly with the same sprayformulation, if different nozzles or operating conditions areused (Combellack, 2004). However, although some gen-eralisations can be made, we are still a long way from beingable to predict appropriate combinations accurately. Inparticular, the characteristics, susceptibility, receptivity, orwhatever term may be applied, of the target, whether plantor other organism, is still poorly understood and may stillbe the dominant variable in the overall efficacy processes.An illustration of possible overall spray efficiency for an

herbicide application is given in Table 1. This is a situationwhere the compounding effects of poor to good efficiencyin individual steps of the deposition:plant retention:upta-ke:translocation interactions will give a very variablesystem efficiency. When individual steps are optimised,the amount received by the target site can be over ahundred and 50-fold more than the worst case.The study of spray formulation efficacy has evolved

from empirical tests and field trials, through laboratoryscreening methods, to more detailed fundamental studiesfor each of the specific steps. Pesticide efficacy can bedefined, and the dose received by the target site ororganism modelled as a function of deposition, retention,

ARTICLE IN PRESS

Table 1

Representative spray application efficiency, system efficiency and off-target loads for an herbicide application

Spray efficacy processes Process efficiency (%) System efficiency (%) Off target component (%)

Depositiona 80–95 80–95 20–5

Retentionb 10–100 8–95 92–5

Uptakec 30–80 2.4–76 97.6–24

Translocationd 10–50 0.24–38 99.6–62

aAmount deposited within target area.bAmount captured by plant.cAmount of retained material taken up into plant foliage.dAmount of absorbed material translocated from absorption site.

J.A. Zabkiewicz / Crop Protection 26 (2007) 312–319 313

uptake, and translocation (Zabkiewicz, 2003). Most workhas been done with herbicides as they involve all the factorsin the spray efficacy processes and account for over half ofall agrichemicals used. Where fungicides or insecticides areconcerned, good plant surface distribution and persistence(due to a lack of systemic compounds), are the predomi-nant requirements. The behaviour and performance ofspray formulations is greatly influenced by adjuvants,either by affecting flow rate, droplet size, dynamic surfacetension, spreading on and wetting of foliage, or byinfluencing uptake and translocation (in plants). A lackof coherent information on the interactions of sprayformulations with fungal and insect organisms precludesany substantive discussion of the factors involved in thesesystems.

This overview will attempt to consider the use of liquidsprays for different applications, their current limitations,and future requirements needed to increase overall efficacyand minimise eco-system loading. It can be done best byconsidering the characteristics and limitations in each ofthe steps of deposition, retention, uptake and transloca-tion.

2. Deposition

Considerable attention has been given to the physicalprocess of spray deposition, and its off-site component,‘‘drift’’. Are the principles applied to aerial application ofsprays relevant to ground-based sprays or to air-blastapplications in orchards? The common denominators aredroplet size, release height and meteorological effects. It isinevitable that aerial sprays will be released at greaterheights above a crop canopy than for ground-basedsystems. Since fall distance is also greater, dropletevaporation can be an issue, leading to potentially moredrift. In this case the addition of anti-evaporant or anti-drift agents (either to retain or increase droplet size) hasbeen exploited. A clear case of formulation technologyused to decrease drift and increase on-site deposition.Access to real-time meteorological information could easilyprevent application when the wind direction is inappropri-ate. Easy to say but difficult to implement or justifyeconomically. The ultimate solution is real-time interactivespray formulation production and application, based on

operating conditions and location of the applicator withinthe spray zone. Smart systems such as these have beensuggested and are under development for ground-basedapplication systems (Ganzelmeier, 2005). To date, im-provements in aerial application have tended to be focusedon the development of spray deposition models, whichhave been very useful in pre-application depositionoptimisation. All operations are at the mercy of meteor-ological conditions, so with such a system, actual deposi-tion profiles can be calculated post-application, using realconditions.In contrast, ground-based application systems have used

a more prescriptive approach, than through optimisationof deposition. However, this may be changing, as concernsdevelop over eco-system loading, and off-site movementinto irrigation ditches or waterways. It has been pointedout that nozzle selection and operating parameters, as wellas formulation factors, need to be tailored to specificapplications (Miller et al., 2001). In the case of groundapplications, drift effects are of the order of a few metres,compared with the hundreds of metres more likely withaerial applications. However, short-range drift may in turngenerate a higher pesticide deposit per unit area, and ifclose to water bodies, exceed permitted levels. One optionis that spraying at field edges should be through nozzlesthat produce courser droplets. This could again be pre-programmed or under the control of the operator. Thoughthis would improve deposition on-site, it may not do so on-target, as it is well known that there is less retention oflarger drops by (hard to wet) plant foliage. It is possiblethat a deliberate change in formulation during application,by the addition of adjuvants, could be a solution. In thecase of systems used on certain fruit crops, shrouded unitswith capture and recycling of the spray solution can reduceboth drift and ground deposition (Ganzelmeier, 1999).

3. Adhesion, retention, re-distribution

There has been considerable focus and study of actualspray retention by crops throughout their growing season,and this approach can quantify soil deposition as well asplant retention (Gyldenkaerne et al., 1999; van de Zandeet al., 2003). Though such empirical field determinationsmay be possible for a range of arable crops, it is unlikely to

ARTICLE IN PRESSJ.A. Zabkiewicz / Crop Protection 26 (2007) 312–319314

be feasible for all crops or weeds in all situations.Furthermore, if pests or diseases are involved, targeteddistribution and placement of the pesticide on or within theplant canopy is even more important. To enhancebiological efficacy in such cases, correct formulationproperties are even more important.

It is well known that droplets of between 50 and 250 mmdiameter will adhere to plant surfaces, but larger dropletsmay rebound or shatter (Hartley and Brunskill, 1958).However, smaller droplets are more prone to drift, so wehave the classic conundrum of how to achieve maximumon-target adhesion/retention, with minimum off-site drift.The situation is complicated by different crops havingdifferent characteristics or pests. The surfaces of targetscan vary widely, from ‘‘easy to wet’’ to almost totallyreflective (with conventional formulations). For example,‘‘wettability’’ (as measured by the contact angle of 20%acetone:water droplets: Forster and Zabkiewicz, 2001) ofupper and lower leaf surfaces and fruit varies from easy tohard to wet (Table 2). This implies not only that adhesion/retention will vary, but also droplet spread, which mayhave considerable significance for pest or disease control.

Since adhesion and retention are dependent on physicalprocesses, they also lend themselves to modelling ap-proaches. A model for droplet primary adhesion (i.e. onfirst impact) has been developed (Forster et al., 2005) andidentifies the main influences as leaf surface character,droplet velocity and spray solution dynamic surfacetension. Similar plant and spray solution parameters areinvolved in retention models (Forster et al., 2004a;

Table 2

Variation in wettablity of plant leaves and fruit surfaces using 20% acetone:w

Species Contact angle (1)

Variety Uppe

Grape Cabarnet sauvignon 93

Apple Pacific Rose 64

Avocado Hass 64

Kiwifruit Hayward 79

aAbstracted from Gaskin et al. (2005).

y = 0.6265x + 0.1837

R2 = 0.9891

0

0.5

1

1.5

2

2.5

3

3.5

0 1 2 3 4 5

Pla

nt r

eten

tion

Surface Deposition µl cm-2

µl cm-2

Fig. 1. Retention of the same series of spray formulations by potato and oni

(Forster et al., 2004a). Dashed line represents 100% retention.

Grayson et al., 1993) but such models are for specific croptypes and plant phenology (which still requires bettercharacterisation), so a universal model does not yet exist.The influence of leaf surface wettability and plantphenology is illustrated in Fig. 1. The same series of sprayformulations were applied to potatoes and onions, twovery dissimilar crops (easy and hard to wet, as well ashaving horizontal and vertical foliage). Retention bypotato foliage followed a simple linear relationship,regardless of spray volume or added adjuvant. On onions,retention did not follow any simple relationship, with poorretention at higher volumes or with less adjuvant. In thissituation, the spraying parameters must be tailored to theplant’s phenology, which is obviously predominant, andwhich is inadequately described in current models.Such retention models are also limited because they do

not include droplet velocities. In the case of field cropsprayers or air-assisted systems, droplet velocities onimpact may be higher than terminal, so primary adhesionand retention efficiencies will be reduced. A practicalsolution which has met with some success, is the additionof ‘‘stickers’’, which are generally high molecular weightadjuvants, to prevent droplet rebound. Although providingbetter adhesion, such additives may not be very usefulwhere good foliar coverage is essential, as they do notspread after adhesion, so need to be applied in smalldroplets. An alternative approach may be the oppositestrategy, which is the use of additives to make largerdroplets shatter on impact with the leaf, and create smallerdroplets in situ within the canopy. This would be an elegant

ater contact angle measurementsa

r leaf Lower leaf Fruit

132 133

64 100

140 86

120 Not available

y = 0.0627x + 0.7612

R2 = 0.0238

0 1 2 3 4 5

Surface Deposition µl cm-2

0

0.5

1

1.5

2

2.5

3

3.5

Pla

nt r

eten

tion

µl cm-2

on plants showing influence of plant leaf wettability and plant phenology

ARTICLE IN PRESSJ.A. Zabkiewicz / Crop Protection 26 (2007) 312–319 315

solution to spray drift but may only be applicable tocertain canopy types. These adhesion enhancing adjuvants(such as organosilicones which provide low surface tensionsolutions) may give not only good adhesion/retention, butby their nature also cause droplet spread over surfaces,including those of insects and fungal bodies. Where topicalplacement and good coverage is paramount, these adju-vants are being used with insecticide and fungicideformulations, particularly on difficult-to-wet plant surfaces(Gaskin et al., 2001). Though adhesion and retentionproperties can be modified substantially by adjuvants, theirinfluence on subsequent performance must also be takeninto account, in particular their effect on uptake andtranslocation.

4. Uptake

In the case of systemic pesticides, if there is little or nouptake, then regardless of the efficiency of deposition andretention, this will have been to no avail. Obviouslyenvironmental factors, such as rain (causing wash-off),wind and relative humidity (affecting droplet drying), willmaterially affect pesticide uptake. However, a major factorin all cases is the plant leaf itself, including its surface andcuticle. One of the main functions of spray adjuvants is toovercome or minimise the effect of leaf waxes and thecuticular barrier. Appropriate adjuvants will substantiallyaid droplet spread, redistribution and leaf coverage, butcan also enhance uptake of the active ingredient. This mayappear obvious, but it is a difficult task for a number ofreasons.

Mass flow of spray solution into leaves through theirstomata was postulated several decades ago (Schonherrand Bukovac, 1972) and demonstrated with foliar fertilizertreatments involving silicone surfactants around the sametime (Neumann and Prinz, 1974) but not put into wider use

Fig. 2. Illustration of droplet spread effects on Chenopodium album with

monododecylether (1.2mm2) or trisiloxane ethoxylate (31.2mm2).

until the 1980s (Stevens and Zabkiewicz, 1988) through theintroduction of the Silwet L-77 organosilicone surfactant.This mechanism, and the properties of sprays required toachieve this effect, have been studied extensively since then(Stevens et al., 1992; Stevens, 1993), and used veryeffectively for certain weed types. However, this mechan-ism is not effective universally for the reason that not allplants have stomata on upper leaf surfaces nor are they ofthe necessary dimensions (Schonherr and Bukovac, 1972).Furthermore, under stress and at night, stomata close,making this mechanism inapplicable.Cuticular uptake (diffusion of the chemical directly

through the cuticle) of pesticides has been studied for alonger time, but progress in understanding the factorscontrolling this mechanism has been much slower. Atten-tion was focussed on relating the physical properties of theformulation to percentage uptake (Stevens and Bukovac,1987; Stock et al., 1993), with less focus on the interactionsof adjuvant(s) in a spray formulation, or the importance ofplant cuticle and epidermal layers. Just as it is essential toconsider application and formulation together in the caseof adhesion and retention by plants, so the interactions ofplant, active and adjuvant must all be considered in theuptake process.The fundamental mechanism of uptake has been

considered (Riederer and Markstadter, 1996; Schonherret al., 1999; Kirkwood, 1999), with most attention givento the epicuticular lipids and their role in modifying activeingredient diffusion through cuticles. However, thereis a much simpler effect on the leaf surface that needs tobe considered first. If a spray formulation containsadjuvants that cause droplet spread on a leaf surface(Fig. 2), this will in effect lower the mass of active per unitarea, without any change in concentration until the spraysolution begins to evaporate. In any case, there will be a‘‘solution residue’’, where the concentration of the active is

different spray formulations. 2-Deoxyglucose with hexaethylene glycol

ARTICLE IN PRESSJ.A. Zabkiewicz / Crop Protection 26 (2007) 312–319316

many times more than in the starting spray solution(Zabkiewicz, 2003).

It has been found (Forster et al., 2004b) that thissolution residue or ‘‘initial dose’’ (ID) can be related to themass uptake of xenobiotics. This relationship has beenvalidated with a wide range of formulations and plants,representing typical field rates and formulations (Forsteret al., 2006). An illustration of such a relationship is givenin Fig. 3 for 2,4-D acid in the presence of various adjuvantsinto Chenopodium album, Hedera helix and Stephanotis

floribunda, where very similar trends are seen for eachspecies. Uptake per unit area at 24 h can be represented bythe relationship: Uptake ¼ a[ID]b, where a and b areconstants specific to the active on these species.

Although the cuticle is not a homogeneous membrane,the diffusion of substances through it may be described byFick’s first law (Price, 1982), where the flux is the amountof solute that diffuses through a unit area per unit of time(i.e. Mass/Area�Time). The flux is proportional to theconcentration gradient and the diffusion coefficient of thexenobiotic. In reality the equation has to include furtherinteractions (Price, 1982), as the solute has to initiallypartition between the external solution and the membrane(partition coefficient between aqueous and lipophilicphases). Other studies (Schonherr and Baur, 1996), usingisolated leaf cuticles, considered transport through thecuticle to be a three stage mechanism. Namely absorptioninto the cuticle, diffusion through the cuticle and finallydesorption from the cuticle into the internal leaf cells.

Schonherr and co-workers (Schonherr et al., 1999) haveidentified that the principal factors affecting uptake rates are:

Solute mobility, which is affected by temperature, solutemolar volumes, and cuticular wax composition.Limiting skin, or the limiting skin tortuosity which is thelength of the diffusion path through the ‘‘limiting skin’’.

0.0001

0.001

0.01

0.1

1

10

100

0.0001 0.01 1 100

Upt

ake

Initial Dose Applied

nmole mm-2

(nmole mm-2)

Fig. 3. 2,4-D mass uptake in the presence of polyethylene glycol

monododecyl ether C12EO3 (open symbols) or a trisiloxane ethoxylate

with mean EO of 7.5 (filled symbols) into Chenopodium album (n,m),

Hedera helix (J,K) and Stephanotis floribunda (&,’) respectively; (- - -)

maximum uptake line, representing 100% uptake over the initial dose

range.

This is only a proportion of, and not the entire cuticle,and is influenced by the size and orientation of thecuticular wax crystals.Driving force, which is affected by the starting andcontinuing concentration of a.i. in the ‘‘solution’’ on thecuticle surface, in the cuticular layers, and in theepidermal cell wall.

Overall, in simple terms,

uptake ¼ solute mobility� cuticle tortuosity

� driving force:

Adjuvants have been classified into ‘‘accelerator’’ and‘‘passive’’ categories, and it is probable that they will affecteach of these terms. In the case of solute mobility, bymodifying xenobiotic solubility in the cuticular waxes; forcuticle tortuosity, by modifying (e.g. swelling) the cuticleand changing its properties; and by modifying the IDthrough droplet spread and hence the concentrationgradient controlling the driving force. The results ofForster et al. (2004b) would indicate that the driving force(related to the ID) is dominant in typical plant systemsusing accelerator adjuvants and hydrophilic to moderatelylipophilic xenobiotics. However, it is unlikely to bepredominant in all situations and all plant species. There-fore, a much better understanding of plant leaf cuticularstructures, as well as structure activity relationships foradjuvants, is still required to progress to a successfulquantitative model of uptake.

5. Translocation

Translocation has received the least attention, thoughlong distance transport has been well studied and reviewed(Price, 1976; Coupland, 1988). In the case of foliar appliedpesticides, it is known that lipophilicity is important, asthese compounds have to cross hydrophobic membranes orstructures other than the cuticle proper. Hydrophilicmolecules are readily transported in either the phloem orthe xylem, though their initial movements through thecuticle, epidermal cells and into the mesophyll are not wellunderstood (Devine and Hall, 1990). The presence ofseparate ‘‘hydrophilic’’ and ‘‘lipophilic’’ pathways as partof the uptake process, may in turn determine the efficiencyof the subsequent translocation pathway, but it isalso difficult to define when uptake becomes translocation(Fig. 4).Previous attempts have been made to relate herbicide

translocation to the concentration of a.i. per droplet, butgave conflicting results, depending on herbicide and con-centration, plant species and formulation differences (Wolfet al., 1992; Liu et al., 1996). Studies have also identifiedthat localised contact phytotoxicity due to formulation orconcentration of specific a.i.s can reduce translocation(Wolf et al., 1992; Forster et al., 1997). It is clear that‘‘long-term’’ translocation will be affected by plant growth

ARTICLE IN PRESS

Fig. 4. Representation of differing trans-cuticular pathways and sub-

sequent apoplastic (polar) and symplastic (non-polar) pathways (adapted

from Steurbaut et al., 1989).

0.00001

0.0001

0.001

0.01

0.1

1

10

100

0.00001 0.001 0.1 10 1000

Mass absorbed nmol mm-2

Mas

s tr

ansl

ocat

ed

nmol mm-2

100% line

Fig. 5. 2-Deoxyglucose mass translocation vs. mass absorbed, in

Chenopodium album, in the presence of different surfactants. Surfactants

are polyethylene glycol monododecyl ethers C12EO3 (’), C12EO6 (K),

C12EO10, (m) and a trisiloxane ethoxylate with mean EO of 7.5 (J); (- - -)

maximum uptake line, representing 100% uptake over the initial dose

range (Zabkiewicz et al., 2004).

0.1

1

10

100

0.1 1 10 100

Mas

s tr

ansl

ocat

ed

nmol mm-2

Mass absorbed nmol mm-2

100% line

Fig. 6. Glyphosate mass uptake into Chenopodium album, vs. mass

absorbed, from Roundup Ultra (m) and Touchdown (’) commercial

formulations; (- - -) maximum uptake line, representing 100% uptake over

the initial dose range (Zabkiewicz et al., 2004).

J.A. Zabkiewicz / Crop Protection 26 (2007) 312–319 317

stage and environmental conditions, but ‘‘short-term’’translocation (over hours rather than days) may haveother rate limiting or modifying features.

Adjuvants are known to facilitate cuticular ‘‘transport’’(foliar uptake) but are not thought to play any significantpart in further short or long-distance translocationprocesses. However, in theory, if adjuvants could reachthe cellular plasmalemma, then they could affect the initialstage of the sub-cuticular transport process (Fig. 4). Therecent use of mass or molar relationships, instead ofpercentages, for xenobiotic uptake into plants fromdiffering formulations (Forster et al., 2004b), may be ameans of elucidating some of the interactions amongactives, adjuvants and plants.

It has been demonstrated (Zabkiewicz et al., 2004) that arelationship similar to that developed for mass uptake, fortranslocation vs. mass absorbed (which is related to the IDapplied) can also apply to the translocation process inmodel systems (Fig. 5) but less well with complexformulations (Fig. 6). The mass uptake relationship is anexpression of the driving force component in Fick’s law ofdiffusion. The fact that a similar relationship has beendemonstrated for translocation with these polar molecules,implies that short distance, or short-term translocation isalso controlled by a concentration gradient effect. Asshown in Fig. 5, there is no difference, either in slope ormolar amount, among the four formulations, indicatingthat once taken up, 2-deoxyglucose translocation is notaffected by these adjuvants. Either they do not penetratefar into the leaf tissues, or they do not have any effect oncell permeability, at these concentrations.

In contrast (Fig. 6) the commercial glyphosate formula-tions show considerable interaction, if the reduction inboth mass uptake and translocation at higher initialdosages, is a correct indication. Previous mass uptakeresults with model glyphosate formulations into otherplant species (Zabkiewicz and Forster, 2001) have alsoshown a divergence from linearity at lower initial dosages

with different adjuvants. This manner of presenting massuptake and translocation has the potential to identifyanomalous behaviour or complex interactions.The advantage of the mass relationship is that it can

provide information similar to that used for drug deliverydose prescriptions. Knowing the mass uptake, an estima-tion can be made of the mass translocated in specificsystems; with subsequent studies on the influence ofphysiological and environmental influences, appropriate‘‘dosages’’ could be applied at specific growth stages orconditions. It can also be used to estimate the change inoverall efficacy of spray formulations, as to date no genericquantitative relationship has been identified.

ARTICLE IN PRESSJ.A. Zabkiewicz / Crop Protection 26 (2007) 312–319318

6. Conclusions

The future focus in the use of pesticides should be ondelivery of the active to the target site, which is not thetotal planted area, but the plant(s) within it, usingappropriate formulations to control spray retention,uptake and translocation. Due to the complexity of thesystem, it is inevitable that this can only be completedthrough the development of computer-based decisionsupport systems (DSS). This in turn requires the develop-ment of models for each of the processes. Depositioninto target areas can be modelled, though this approachis not used much by most of the primary sectors. Therequirement to place deposits onto specific parts of aplant, or all over a plant, needs to be addressed througha much better approach to modelling retention (insteadof empirical data sets) which in turn requires a betterdescription of plant development and (foliar) sur-face characteristics. The principal factors controllingfoliar uptake appear to be solute mobility in the cuticle,cuticle tortuosity and solute-driving force. Apart fromthe last factor, the other two are poorly understoodand not easy to measure. Whole plant translocationmodels exist, but it is the short-term or short-distancecontrols that need to be elucidated and related to the initialmass uptake.

More complex models of uptake or total pesticideefficacy are being developed and show promise (Satchiviet al., 2000a, b; Lamb et al., 2001), but their universalapplicability and acceptance will not occur until the inputparameters have been substantially simplified or can bemeasured easily. At present, as with all models, theirgreatest value may lie in identifying the gaps in ourknowledge, so that we focus on the most importantrequirements

The integration of spray deposit profiles within geo-graphic information systems can provide information thatcan be subsequently related to pest or weed competitionlevels, crop productivity and cumulative residue profiles.The development of a comprehensive DSS has the furtherbenefits of providing tailored sub-sets of information thatcan be used by individual operators, for operationaloptimisation, such as selection of appropriate pesticidesor application technology and maximising on-targetdeposits. This DSS can in turn become part of anagrichemical management system, which can be appliedon a local, regional or national scale and related to long-term pesticide eco-system interactions.

Acknowledgements

This work was supported by the New Zealand Founda-tion for Science and Technology. Thanks are due to W.A.Forster and R.E. Gaskin for constructive advice and use ofdata, and M. Haslett for assistance with the preparation ofthe figures for this manuscript.

References

Combellack, J.H., 2004. Interaction of nozzles with spray fluids on droplet

drift. Plant Prot. Q. 19, 119–124.

Coupland, D., 1988. Factors affecting the phloem translocation of foliage

applied herbicides. In: Aitken R.K., Clifford D.R. (Eds.). British Plant

Growth Regulation Group Monograph 18, pp. 85–112.

Devine, M.D., Hall, L.M., 1990. Implications of sucrose transport

mechanisms for the translocation of herbicides. Weed Sci. 38, 299–304.

Forster, W.A., Zabkiewicz, J.A., Murray, R.J., Zedaker, S.M., 1997.

Contact phytotoxicity of triclopyr formulations on three plant species

in relation to their uptake and translocation. In: O’Callaghan, M.

(Ed.), Proceedings of the 50th New Zealand Plant Protection

Conference, vol. 50. Lincoln, New Zealand, pp. 125–128.

Forster, W.A., Zabkiewicz, J.A., 2001. Improved method for leaf surface

roughness characterisation. In: De Ruiter, H. (Ed.), Proceedings of

the Sixth International Symposium on Adjuvants for Agrochemicals,

pp. 113–118.

Forster, W.A., Steele, K.D., Gaskin, R.E., Zabkiewicz, J.A., 2004a. Spray

retention models for vegetable crops: preliminary investigation. In:

Zydenbos, S.M. (Ed.), New Zealand Plant Protection, vol. 57.

Hamilton, New Zealand, pp. 125–128.

Forster, W.A., Zabkiewicz, J.A., Riederer, M., 2004b. Mechanisms of

cuticular uptake of xenobiotics into living plants: 1. Influence of

xenobiotic dose on the uptake of three model compounds, applied in

the absence and presence of surfactants into Chenopodium album,

Hedera helix and Stephanotis floribunda leaves. Pest. Manage. Sci. 60,

1105–1113.

Forster, W.A., Kimberley, M., Zabkiewicz, J.A., 2005. A universal spray

droplet adhesion model. Transactions of ASAE 48, 1321–1330.

Forster, W.A., Zabkiewicz, J.A., Riederer, M., 2006. Cuticular uptake of

xenobiotics into living plants: 2. Influence of xenobiotic dose on the

uptake of bentazone, epoxiconazole and pyraclostrobin, applied in the

presence of various surfactants into Chenopodium album, Sinapis alba

and Triticum aestivum leaves. Pest. Manage. Sci. 62, 664–672.

Ganzelmeier, H., 1999. Plant protection—current state of technique and

innovations. In: Brooks, G.T., Roberts, T.R. (Eds.), Pesticide

Chemistry and Bioscience. Proceedings of the Ninth IUPAC Congress.

Royal Society of Chemistry, Cambridge, pp. 100–119.

Ganzelmeier, H., 2005. GIS-based application of plant protection

products. Ann. Rev. Agricultural Engineering 4, 245–256.

Gaskin, R.E., Elliott, G.S., Munro, J.P., Murray, R., 2001. Improving

spray performance on onion crops with novel organosilicone adjuvant

blends. In: De Ruiter, H. (Ed.), Proceedings of the Sixth International

Symposium on Adjuvants for Agrochemicals, pp. 327–332.

Gaskin, R.E., Steele, K.D., Forster, W.A., 2005. Characterising plant

surfaces for spray adhesion and retention. In: Zydenbos, S.M. (Ed.),

New Zealand Plant Protection, vol. 58, pp. 179–183.

Grayson, B.T., Pack, S.E., Edwards, D., Webb, J.D., 1993. Assessment of

a mathematical model to predict spray deposition under laboratory

track sprayer conditions. II: Examination with further plant species

and diluted formulations. Pestic. Sci. 37, 133–140.

Gyldenkaerne, S., Secher, B.J.M., Nordbo, E., 1999. Ground deposit of

pesticides in relation to the cereal canopy density. Pestic. Sci. 55,

1210–1216.

Hall, F.R., 2004. The importance of spray drift management around the

world. In: Ramsay, C., Hewitt, A., Thistle, H., Hoffman, C., Wolf, R.,

Wolf, T. (Eds.), Proceedings of the International Conference on

Pesticide Application for Drift Management. Washington State

University, Pullman, WA, USA, pp. 20–27.

Hartley, G.S., Brunskill, R.T., 1958. Reflection of water droplets from

surfaces. In: Danielli, J.F., Parkhurst, K.G.A., Riddiford, A.C. (Eds.),

Surface Phenomena in Chemistry and Biology. Pergamon Press, New

York, pp. 214–223.

Kirkwood, R.C., 1999. Recent developments in our understanding of the

plant cuticle as a barrier to the foliar uptake of pesticides. Pestic. Sci.

37, 69–77.

ARTICLE IN PRESSJ.A. Zabkiewicz / Crop Protection 26 (2007) 312–319 319

Lamb, S., Bean, M., Chapman, P., 2001. Foliar uptake of pesticides by

mathematical modelling: a tool for influencing adjuvant selection. In:

de Ruiter, H. (Ed.), Proceedings of the Sixth International Symposium

on Adjuvants for Agrochemicals. ISAA 2001 Foundation, The

Netherlands, pp. 106–112.

Liu, S.H., Campbell, R.A., Studens, J.A., Wagner, R.G., 1996. Absorp-

tion and translocation of glyphosate in aspen (Populus tremuloides) as

influenced by droplet size, droplet number and herbicide concentra-

tion. Weed Sci. 44, 482–488.

Miller, P.C.H., Lane, A., Wheeler, H., 2001. Matching spray applications

to canopy characteristics in cereal crops. Pestic. Outlook, 100–102.

Neumann, P.M., Prinz, R., 1974. Evaluation of surfactants for use in the

spray treatments of iron chlorosis in citrus trees. J. Sci. Food Agric. 25,

221–226.

Price, C.E., 1976. Penetration and translocation of herbicides and

fungicides in plants. In: McFarlane, N.R. (Ed.), Herbicides and

Fungicides-Factors Affecting their Activity. The Chemical Society

Special Publication 29, London, pp. 42–66.

Price, C.E., 1982. Penetration and translocation of herbicides and

fungicides. In: Cutler, D.F., Alvin, K.L., Price, C.E. (Eds.), The Plant

Cuticle. Academic Press, London, UK, pp. 237–252.

Riederer, M., Markstadter, C., 1996. Cuticular waxes: a critical

assessment of current knowledge. In: Kerstiens, G. (Ed.), Plant

Cuticles—An Integrated Functional Approach. Bios Scientific Publish-

ers, Oxford, UK, pp. 189–200.

Schonherr, J., Baur, P., 1996. Effects of temperature, surfcatants

and other adjuvants on rates of uptake of organic compounds.

In: Kerstiens, G. (Ed.), Plant Cuticles—An Integrated

Functional Approach. Bios Scientific Publishers, Oxford, UK,

pp. 134–154.

Schonherr, J., Baur, P., Buchholz, A., 1999. Modelling foliar penetration:

its role in optimising pesticide delivery. In: Brooks, G.T., Roberst,

T.R. (Eds.), Pesticide Chemistry and Bioscience. Royal Society of

Chemistry, Cambridge, pp. 134–154.

Schonherr, J., Bukovac, M.J., 1972. Penetration of stomata by liquids.

Dependence on surface tension, wettability and stomatal morphology.

Plant Physiol. 49, 813–819.

Satchivi, N.M., Stoller, E.W., Wax, L.M., Briskin, D.P., 2000a. A

nonlinear dynamic simulation model for xenobiotic transport and

whole plant allocation following foliar application I. Conceptual

foundation for model development. Pestic. Biochem. Physiol. 68,

67–84.

Satchivi, N.M., Stoller, E.W., Wax, L.M., Briskin, D.P., 2000b. A

nonlinear dynamic simulation model for xenobiotic transport and

whole plant allocation following foliar application II. Model valida-

tion. Pestic. Biochem. Physiol. 68, 85–95.

Steurbaut, W., Melkebeke, G., DeJonckheere, W., 1989. The influence of

nonionic surfactants on the penetration and transport of systemic

fungicides in plants. In: Chow, P.N.P., Grant, C.A., Hinshalwood,

A.M., Simundsson, E. (Eds.), Adjuvants and Agrochemicals, vol. 1.

CRC Press, Boca Raton, USA, pp. 93–103.

Stevens, P.J.G., 1993. Organosilicone surfactants as adjuvants for

agrochemicals. Pestic. Sci. 38, 103–122.

Stevens, P.J.G., Bukovac, M.J., 1987. Studies on octylphenoxy surfac-

tants. Part 1: effects of oxyethylene content on properties of potential

relevance to foliar absorption. Pestic. Sci. 20, 19–35.

Stevens, P.J.G., Gaskin, R.E., Hong, S.O., Zabkiewicz, J.A., 1992. In:

Foy, C.L. (Ed.), Adjuvants for Agrichemicals. CRC Press, Boca

Raton, USA, pp. 385–398.

Stevens, P.J.G., Zabkiewicz, J.A., 1988. Effect of surfactants on foliar

uptake: interactions with species, chemicals and concentrations. In:

Proceedings of the EWRS Symposium: Factors Affecting Herbicidal

Activity and Selectivity, Wageningen, pp. 145–150.

Stock, D., Holloway, P.J., Grayson, B.T., Whitehouse, P., 1993.

Development of a predictive uptake model to rationalise selection of

polyoxyethylene surfactant adjuvants for foliage applied agrochem-

icals. Pestic. Sci. 37, 233–245.

Van de Zande, J.C., Parkin, C.S., Gilbert, A.J., 2003. Application

Technologies. In: Wilson, M. (Ed.), Optimising Pesticide Use. Wiley,

New York, pp. 23–44.

Wolf, T.M., Caldwell, B.C., McIntyre, G.I., Hsiao, A.I., 1992. Effect of

droplet size and herbicide concentration on absorption and transloca-

tion of 2,4-D in oriental mustard (Sisymbrium orientale). Weed Sci. 40,

568–575.

Zabkiewicz, J.A., 2003. Foliar interactions and uptake of agrichemical

formulations—present limits and future potential. In: Voss, G.,

Ramos, G. (Eds.), Chemistry of Crop Protection. Wiley VCH,

Weinheim, pp. 237–251.

Zabkiewicz, J.A., Forster, W.A., 2001. Percentage uptake to molar flux-

approaches to modelling in vivo cuticular pesticide uptake. In: de

Ruiter, H. (Ed.), Proceedings of the Sixth International Symposium on

Adjuvants for Agrochemicals. ISAA 2001 Foundation, Amsterdam,

pp. 19–28.

Zabkiewicz, J.A., Forster, W.A., Gaskin, R.E., Hofstee, M., 2004. Foliar

uptake and translocation relationships for polar xenobiotics. In:

North, M. (Ed.), Proceedings of the Seventh International Symposium

on Adjuvants for Agrochemicals. ISAA 2001 Foundation, Cape Town,

pp. 242–247.