You are on page 1of 13

Journal of Experimental Botany, Vol. 57, No. 11, pp.

2501–2513, 2006
doi:10.1093/jxb/erl023 Advance Access publication 7 July, 2006

FOCUS PAPER

Characterization of the diffusion of non-electrolytes across


plant cuticles: properties of the lipophilic pathway

Anke Buchholz*
Syngenta Crop Protection AG, Research Biology, CH-4332 Stein, Switzerland

Received 25 November 2005; Accepted 10 April 2006

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


Abstract Introduction
Systemic crop protection products are commonly Foliar uptake is a complex process determined by several
sprayed onto foliage, whereupon the active substances mutually affecting parameters. General characteristics—
must penetrate into the leaves in order to become required for the optimization of foliar uptake—can only be
biologically active. Penetration of the plant cuticle is obtained when interdependencies have been excluded. Ex-
the rate-limiting step. The diffusion of organic non- tensive work, focusing on the diffusion of non-electrolytes
electrolytes within cuticles is a purely physical pro- across plant cuticles, has been done, which this paper
cess that can be described and analysed in the same attempts to review. The relevant characterization of cu-
way as is done for diffusion in synthetic polymer ticle penetration for polar compounds is given separately
membranes. Solute mobilities in cuticles vary consid- by Schönherr (2006). After a general introduction on
erably between plant species. For a given species the function and character of plant cuticles, the applied
they decrease with increasing solute size, and this methodology is presented. Parameters affecting plant
size selectivity holds for all of the plant species in- cuticle penetration are displayed individually so that finally
vestigated so far. Wax extraction from leaf cuticles the complex scenario of foliar uptake can be outlined in
increases the mobility of solutes tremendously, but a simplified way.
size selectivity is not affected. Furthermore, diffusion
within plant cuticles is extremely temperature depen-
dent. An analogous increase in solute mobility can Function of plant cuticles
be achieved by using accelerators, which enhance
All aerial surfaces of the primary parts of terrestrial
the fluidity of the polymer matrix and of the waxes.
higher plants are covered with a cuticle, the function of
The effects of temperature and plasticizers on the dif-
which is to prevent uncontrolled water loss (Riederer
fusion of non-electrolytes in wax and the cutin matrix
and Schreiber, 2001). The plant cuticle is not only exposed
have been used to characterize the nature of the lipo-
to abiotic factors like sunlight, wind, rain, etc. but it
philic pathway. The ‘free volume’ theory can be used
also interacts with microbes, fungi, and insects (Juniper,
to explain the influence of the size and shape of
1995; Kolattukudy et al., 1995; Kerstiens, 1996; Schreiber
the solute, and its dependence on temperature. The
et al., 2004). Furthermore, the plant cuticle is the initial
physico-chemical nature of the diffusion pathway has
contact point between an agrochemical and the plant and,
been shown, by thermodynamic analysis, to be ident-
where uptake into plant tissues is required for biological
ical for a wide range of solute lipophilicities. This ap-
activity, it is the main barrier to penetration.
proach also explains the mode of action and the
intrinsic activity of plasticizers.
Morphology, chemistry, and structure of
Key words: Cuticular membrane, diffusion, foliar uptake, leaf
plant cuticles
surface, penetration, solute mobility.
Cuticles are not simple homogeneous structures like syn-
thetic polymer films. Various structural types occur in

* E-mail: anke.buchholz@syngenta.com
Abbreviations: CM, cuticular membrane; k*, solute mobility; MX, matrix membrane; UDOS, unilateral desorption of the outer surface; Vx, molar volume.

ª The Author [2006]. Published by Oxford University Press [on behalf of the Society for Experimental Biology]. All rights reserved.
For Permissions, please e-mail: journals.permissions@oxfordjournals.org
2502 Buchholz
gymnosperms and angiosperms, which were categorized waxes (Bianchi, 1995). However, the permeability of this
by Holloway (1982) into six morphological types. Inves- barrier is independent of the overall thickness of the cuti-
tigators of the cuticular ultrastructure of different plant cle (Becker et al., 1986; Knoche et al., 2000). The intra-
species have used different terms to describe these layered cuticular waxes are predominantly located in the outer
structures. Wattendorf and Holloway (1980) provided a layers of the cuticle where they form the transport-limiting
comparative overview of the different terminologies used. layer or skin (Schönherr and Riederer, 1988). This wax
The following generalization as given by Jeffree (1986) layer acts as a barrier to both diffusion and solubility
(Fig. 1) distinguishes three main zones. The cuticle proper (Shafer and Schönherr, 1985), and it has been established
with epicuticular and intracuticular waxes often has a that the solid and crystalline wax aggregates within this
lamellate layer and typically extends to a thickness of layer determine the transport properties of the plant cuticle
50–150 nm (Jeffree, 1996). The cuticular layer is bonded (Riederer and Schreiber, 1995).
to the periclinal walls of the epidermal cells by a pectin-
rich layer. This pectinaceous layer is equivalent to the

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


middle lamella, with which it is continuous. Enzymatic
hydrolysis of this layer enables the cuticular membranes Penetration across plant cuticles
(Orgell, 1955), which comprise the cuticle proper and the The penetration of non-electrolytes across plant cuticles
cuticular layer, to be isolated. requires three steps: sorption into the cuticular lipids, dif-
The epicuticular wax film is a distinct layer on the fusion across the cuticular membrane, and finally desorp-
surface of the cutin matrix (Jetter et al., 2000). Depending tion into the apoplast of the epidermal cells (Kirkwood,
on their chemical composition, epicuticular waxes may 1999; Schönherr et al., 1999). Rates of penetration de-
form an amorphous film, granules, or crystalline structures pend upon solute mobility in the cuticle and on the driving
of various shapes (Baker, 1982; Jeffree, 1986). This fine- force (product of partition coefficient and concentration
surface structure might be modified with increasing age gradient; see also below). Solute mobility can be meas-
(Bukovac et al., 1979), by environmental stress (Hoad ured using unilateral desorption from the outer surface
et al., 1992), and also by agrochemical spray solutions (UDOS). This method takes advantage of the fact that
(Whitehouse et al., 1982; Tamura et al., 2001). Epicuticu- cuticles are asymmetric membranes (see above) and distin-
lar waxes have a considerable influence on the wettability guishes two functional layers: the limiting skin and the
of plant surfaces, often causing poor retention or spread- sorption compartment (Fig. 2). In the cuticles studied, the
ing of spray droplets (Bukovac et al., 1979; Turunen and thin limiting skin comprised only about 10% of the total
Huttunen, 1990; Holloway, 1994), but they rarely affect
a compound’s rate of penetration into leaves (Baur, 1998).
The penetration barrier (Schönherr and Schmidt, 1979)
is made up of the cuticular membrane, which consists of
the cutin polymer matrix (Holloway, 1993) and associated

Fig. 2. Functional layers (not to scale) of cuticular membranes as


differentiated in the UDOS (unilateral desorption of the outer surface) and
SOFU (simulation of foliar uptake) methods: limiting skin (LS) and
sorption compartment (SC). Occurrence and dimension of cuticular pegs,
and protrusion of the cuticular layer into anticlinal cell walls of epidermal
cells, are species dependent. The arrows denote the direction of diffusion
according to the concentration gradient. For one time, the sorption
Fig. 1. Generalized structure of a plant cuticle (modified according to compartment (UDOS) and, for the other time, the formulation residue
Jeffree, 1986). EW, Epicuticular wax; CP, cuticle proper with lamellate facing the limiting skin (SOFU) serve as donors. With UDOS rate
structure; CL, cuticular layer traversed by cellulose microfibrils; PL, constants of desorption, k*, and with SOFU rate constants of penetration,
pectinaceous layer and middle lamella; CW, cell wall; P, plasmalemma. k, can be determined.
Lipophilic pathway 2503
mass of the cuticular membrane (CM), and both sorption where K is the partition coefficient and Dx is the membrane
capacity and solute mobility were very small. The sorption thickness. The reader is referred to Bauer and Schönherr
compartment underneath the limiting skin amounted to (1992), Schönherr and Baur (1994), and Baur et al. (1996b)
about 90% of the mass of the CM, its sorption capacity for further details.
was high, and solute mobility was much higher than in The benefit of this method is that the contributions
the limiting skin (Schönherr and Baur, 1994). Model com- of mobility and solubility to the rates of foliar uptake
pounds are applied to the morphological inner surface of pesticides can be determined separately, due to the fact
of the CM and are subsequently sorbed in the sorption that the measured rate constants of desorption (k*) are
compartment. When they are homogeneously distributed independent of the partition coefficient. When the donor
within the sorption compartment they are unilaterally de- is applied to the outer surface of the CM and solutes are
sorbed from the morphological outer surface. desorbed from the morphological inner side (SOFU=
These measured solute mobilities are first-order rate simulation of foliar uptake), rates of desorption also
constants of desorption (k*), which are independent of follow first-order kinetics, but they depend on both solu-

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


the partition coefficient and are directly proportional to bility and mobility (Baur et al., 1997b).
the diffusion coefficient in the waxy outer limiting skins
of cuticles.
Plant species
D = k*3lls 3lsoco ð1Þ
When solute mobility in cuticles is studied using the
where lls3lsoco is the diffusion path length in the limiting UDOS method, astomatous CM must be used. Enzymatic
skin, ls, and sorption compartment, soco. The translation isolation does not affect the transport properties of cuti-
to diffusion coefficients D (m2 s1) is only restricted cles (Kirsch et al., 1997) but it is feasible only with a
since the diffusion path length is not known precisely limited number of species, i.e. those where a continuous
(Schönherr and Baur, 1994). When only the permeability pectinaceous layer is present. These are most often peren-
P is determined, sorption processes are additionally in- nial plants. The range of solute mobilities (e.g. for bifenox)
cluded and cannot be separated. measured for the CMs studied covered three orders of
magnitude (Fig. 3). These different mobilities are species-
P = ðD3KÞ=Dx ð2Þ specific characteristics and do not correlate with either

log k*
-4.5 -5 -5.5 -6 -6.5 -7 -7.5 -8 -8.5
Prunus persica
Tilia cordata
Populus canescens
Juglans regia
Pyrus pyrifolia
P. communis cv. Conference
Prunus armeniaca
Stephanotis floribunda
Malus baccata
Pyrus communis
Pyrus pyrifolia
M. domestica cv. Golden D.
M. domestica cv. Gloster
Citrus grandis
Prunus serotina
Ilex aquifolium
Prunus laurocerasus
Citrus aurantium
Strophanthus gratus
Ginkgo biloba
Melicoccus bijugatus
Hedera helix
Vanilla planifolia
Ilex paraguariensis
Decreasing solute mobility

Fig. 3. Solute mobilities, logk*, of the model compound bifenox (logKCW, 4.44; molar volume, 216 cm3 mol1) measured in 24 plant species (CM) at
25 8C. Error bars denote 95% confidence intervals.
2504 Buchholz
cuticle thickness or the amount of wax embedded within, homologous waxes (1-tetradecanol and 1-octadecanol) re-
and covering, the cuticular matrix. For example, Pyrus vealed very complex mixing characteristics, and suggested
communis, Malus domestica, Ilex aquifolium, and Ilex frequent phase immiscibility (Carreto et al., 2002). There-
paraguariensis exhibited a very similar amount of ex- fore, a mosaic of phase domains had to be expected even
tractable wax (about 24%) but its contribution to the barrier at ambient temperatures.
property of the CM varied by factors of 49 up to 4897 This coexistence of crystalline and fluid or amorphous
(Table 1). No clear relationship has been found to exist phases was determined for the cuticular waxes extracted
between the particular composition of the soluble lipids from several different plant species (Reynhardt and
and the respective cuticular permeability (Haas and Riederer, 1991, 1994; Viougeas et al., 1995; Schreiber
Schönherr, 1979; Riederer and Schneider, 1990; Hauke et al., 1997), and was also demonstrated in situ (Merk
and Schreiber, 1998). Histochemical methods have also et al., 1998). The measurement of the diffusion character-
been used to try to understand the relationship between istics in reconstituted waxes is thought to be a reason-
chemical composition and permeability (Leece, 1976; able approach since the composition, and the proportion of

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


Krüger et al., 1996). crystallites and amorphous phases is not altered during
recrystallization (Schreiber and Schönherr, 1993). A good
correlation between mobilities measured in reconstituted
cuticular waxes and in CMs was found, which confirmed
Cuticular waxes and cutin matrix
viscosity as an important property (Schreiber and Riederer,
Cuticular wax composition is very complex and varies 1996; Kirsch et al., 1997). These investigations revealed
according to species, ontogenetic development, and envir- that the degree of crystallinity and the spatial arrangement
onmental conditions (Bianchi, 1995; von Wettstein- of crystals determined the permeating path through the
Knowles, 1995). The commonest components are n-alkanes amorphous phase, which in turn determined the effective-
and primary alcohols, both of which range from C16 to C36 ness of the diffusion barrier (Schreiber et al., 1997).
(Holloway, 1994). Long alkyl chain compounds exhibit Extracted waxes might, however, crystallize in a differ-
a solid-state polymorphism (liquid!hexagonally packed ent shape (spatial extension) compared with those con-
solid phase!orthorhombically packed solid phase) across fined in a polymer matrix. Due to tortuosity (the path
a temperature scan (Watanabe, 1961). The physical state- taken around the crystalline obstacles) the length of the
of the cuticular wax may contain several phases and diffusion path can be much greater than the thickness of
crystalline forms and is likely, therefore, to be the critical the barrier (Schreiber et al., 1996a; Baur et al., 1999).
feature that determines the transport properties of the cuti- The study of synthetic chemical polymer membranes
cle. Highly ordered crystalline regions are inaccessible to can also help to elucidate the mechanism of diffusion
permeation. The physical state of a mixture is a function across CMs, even though they are, in general, thicker and
of its chemical composition and of physical variables like less complex than plant cuticles. This enables a more de-
temperature, pressure, etc. and is best described using a tailed view of the functional relationships to be obtained.
phase diagram. Phase diagrams for such complex mixtures For instance, clay platelets are incorporated as inert fillers
as cuticular waxes are difficult to obtain and would not into polymers to improve their mechanical and barrier
be very informative due to their complexity. The ex- properties. To make the clay platelets compatible with
amination of a simple binary mixture of two chemically hydrophobic materials, such as polyolefins and waxes,

Table 1. Comparison of solute mobilities (k*; 695% CI) for bifenox within CM and MX of 13 plant species at 25 8C
Extractable wax fraction (mg cm2; weight percentage) and differential factors of both solute mobilities are specified.
Plant species CM log k* 95% CI MX log k* 95% CI Factor Wax (mg cm2) Wax (%)

Populus canescens –5.00 0.07 – – – 0.053 33.9


Pyrus pyrifolia –5.34 0.17 –3.29 0.05 112 0.055 25.5
P. communis cv. Conference –5.43 0.11 –3.74a – 49 0.061 23.7
Stephanotis floribunda –5.68 0.25 –3.73 0.02 89 0.026 10.2
M. domestica cv. Golden Delicious –6.20 0.08 –3.36 0.06 692 0.044 23.7
Ilex aquifolium –6.64 0.18 –4.14 0.12 316 0.162 24.6
Ginkgo biloba –6.97 0.11 –5.54 0.19 27 0.054 13.4
Melicoccus bijugatus –7.04 0.11 –4.81 0.06 170 0.049 8.8
Hedera helix –7.16 0.14 –3.75 0.05 2570 0.022 10.2
Strophanthus gratus –7.18 0.16 –5.34a – 69 0.060 19.5
Vanilla planifolia –7.67 0.11 –6.93 0.06 5 0.061 10.8
Ilex paraguariensis –7.89 0.11 –4.20 0.15 4897 0.114 22.3
Schefflera actinophylla –8.09a – –4.57a – 3311 0.084 29.8
a
Solute mobilities at 25 8C were calculated from Arrhenius graphs (Buchholz, 1998).
Lipophilic pathway 2505
their surfaces are treated with monolayers of specific Marga et al. (2001) specified that the chemical composition
amines. The mechanism by which these organoclays im- determined both the biochemical properties as well as
prove the barrier properties of the film relies on the high the elastic or rigid behaviour.
aspect ratio (relationship between the width and height) Consequently, several structural principles described
of the exfoliated clay platelets to impart a tortuous diffu- for synthetic polymers can be applied to the plant cuticle.
sion path. The tortuosity factor can be as high as several- The plant cuticle can be described as a matrix composed
hundred-fold for impermeable platelets with aspect ratios of highly ordered crystalline and disordered amorphous
of 100–500 and at modest mineral loadings of 5–10 regions, within which the reduced dynamics in the crys-
vol% (Cussler et al., 1988). talline regions of the wax barrier make them practically in-
The tortuosity of nanocomposite materials, however, accessible to permeants (Reynhardt and Riederer, 1991).
is rarely that high (typically 2- to 4-fold) and it has been The cutin polymer not only serves as a matrix but also
shown that the interphase surrounding the clay platelets contributes to the transport characteristics of the cuticle
exerts a profound influence on gas permeation (Chaiko (Santier and Chamel, 1998). Cutin monomers are cross-

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


and Leyva, 2005): The polymer systems, the polyolefins linked by ester, peroxide, and ether bridges (Kolattukudy,
together with the waxes, are semi-crystalline materials 1984), and the presence of large amounts of di- and
consisting of crystalline and amorphous phases. If the hydro- tri-hydroxyfatty acids may affect the degree of cross-
carbon chains are long enough, they will fold in on them- linking. It was demonstrated that the maturation of cutin
selves to form crystallites (10–20 nm). The crystallites led to the addition of covalent bonds which resulted in
are held together by amorphous chain segments that less polarity and reduced permeability (Schmidt et al.,
contribute to the elastic strength of the material. If the 1981). The wax-free matrix membranes (MXs) of several
hydrocarbon chain length is shortened sufficiently, the am- plant species showed a relatively wide range of solute
orphous chain segments become less able to bridge mobilities (Table 1). The applicability of the UDOS
the gap between the crystallites, and the material becomes method also provided evidence for the asymmetry of the
brittle. Gas diffusion can then take place at the interfaces cutin matrix. Chloroform–methanol extraction of CM
between the crystallites. to provide MX not only removes the impermeable wax
Similar interactions were also demonstrated with Hor- crystallites but also the amorphous waxes. Therefore, the
deum wax. Longer chains bridged the amorphous zone elevated solute mobilities in MX are due to the reduced
between two adjacent layers of crystalline material, thus viscosity (absence of amorphous waxes) and the shortened
linking the two layers (Reynhardt and Riederer, 1994). diffusion path (lack of crystalline waxes causing tortuosity).
The relevance of tortuosity in cuticular waxes as an effec-
tive way of reducing the permeabilities of plant cuticles
is discussed by Schreiber (2006). Solute size effects
A significant amount of shrinkage might accompany
wax recrystallization due to the high degree of crystall- Size dependence of mobility was quantified by plotting
inity and the large density difference between the amorph- for each plant species logk* versus molar volume (Vx)
ous melt and the crystal phase. As the freezing points of for the respective sets of model compounds. Characteristic
the individual components in the wax differ significantly molar volumes were calculated according to McGowan
(Chaiko and Leyva, 2005), this shrinkage could lead to and Mellors (1986). These molar volumes represent the
extensive cracking. Nanocomposite materials mixed with volumes at absolute zero temperature and they are closely
polyethylene, however, displayed reduced embrittlement correlated to van der Waals volumes (McGowan, 1969).
at high clay loadings, and the films were quite flexible Solute mobilities in the CMs of numerous plant species
and free of gross structural defects. Within plant cuticles have been measured using linear (Baur et al., 1996b) and
the cutin polymer can be viewed as the analogous elas- cyclic organic model compounds varying in molar volume
tomer ensuring flexibility and avoiding embrittlement. from 99 to 349 cm3 mol1 (Bauer and Schönherr, 1992;
The plant cuticle has to be flexible to follow deforma- Schönherr and Baur, 1994; Buchholz et al., 1998; Baur
tions caused either by fluctuations in turgor pressure or et al., 1999). Solute mobilities (k*) decreased exponent-
by wind stress. Petracek and Bukovac (1995) investigated ially with increasing molar volumes (Vx) and, with all
the rheological properties of tomato fruit cuticles and plant species tested, good linear correlations between
identified the cuticle as a viscoelastic polymer, com- logk* and Vx were found. This dependence could be
prising both elastic (reversible) and plastic (irreversible) described with equations of the type
components. Furthermore, water affected the mechanical
logk* = logk*0  b9Vx ð3Þ
properties, and waxes reduced elasticity and susceptibility
to fracturing. The mechanical properties of cuticles are The y-intercepts (k*0) represent the mobility of a hypo-
not related to thickness but are likely to be altered by thetical molecule having zero molar volume and the
chemical composition (Wiedemann and Neinhuis, 1998). slopes of the graphs (b9) represent the size selectivity of
2506 Buchholz

7
-log k *

5 Hedera helix
-log k * = 5.22 + 0.009 . V x
Malus domestica
4 -log k * = 4.11 + 0.010 . Vx

Populus canescens

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


-log k * = 2.33 + 0.011 . Vx Fig. 5. Exponential hole-size distribution of a polymer matrix as
3
affected by temperature (modified according to Lieb and Stein, 1972).
50 100 150 200 250 300 350 400 (A) The distribution function provides the probability (P) of finding
a free-volume hole of a specific size (v*) at a given temperature (T). v* is
Vx (cm3 . mol-1) the minimum volume hole size into which a molecule can jump. The
diffusion coefficient is considered to be proportional to the probability
Fig. 4. The effect of molar volumes (Vx) of model compounds on their of finding a hole of volume v* or larger. (B) Increasing temperature
mobility (k*) in cuticular membranes of three selected plant species. increases the overall free volume available for migration (shaded area).
Averages of 16–20 cuticular membranes and 95% confidence intervals
are shown.

the barrier. The differences in size selectivity (b9) were The benefit of this relationship for the estimation of
not significant but the y-intercepts ranged from 2.33 the solute mobilities of new compounds at a given temper-
(Populus) to 5.22 (Hedera) (Fig. 4). The latter are pure ature and plant species has been impressively demonstrated
membrane parameters, which are completely indepen- (Buchholz et al., 1998). If the temperature sensitivity of
dent of the properties of solutes (size and lipophilicity), a plant species is known for only two compounds (i.e.
and they can be related to the free volume available for having two respective Arrhenius graphs) then a graph
diffusion in cuticles and cuticular waxes. (The real conver- of b9 versus temperature can be made. With this infor-
gence of this graph with the y-intercept is not known as mation, the size selectivity (b9) for any desired temperature
data describing the mobilities of non-electrolytes having can be derived and consequently the solute mobility for
molar volumes significantly below 100 cm3 mol1 are still any third compound (Vx) is predictable.
unavailable.) An explanation for this (temperature-dependent) size
No significant differences were observed in b9 among discrimination can be derived from the free volume
all the species tested, with the average value of b9 being theory (Coughlin et al., 1991; Duda and Zielinski, 1996).
0.0095 mol cm3 at 25 8C. This equates to a reduction For a successful diffusion step within a polymer the pen-
in solute mobility of 700 times if Vx is increased from etrant needs a hole (a space between polymer chains) di-
100 to 400 cm3 mol1. This size discrimination was found rectly adjacent to its actual position. Enough energy to
to be the same when cuticular waxes were extracted and make the jump is ensured by the motion of the solute
comparable size selectivities were determined for poly- itself, via Brownian motion. Hence it follows that only
mer MXs (Baur et al., 1999). A similar correlation the movement of the adjacent polymer chains limits the
between diffusivity and solute size was also found for solute’s spatial motion. At a given temperature, a polymer
reconstituted waxes (Schreiber and Schönherr, 1993; has a characteristic distribution of these temporary holes.
Schreiber et al., 1996a; Schreiber, 2006). Elevating the temperature causes an increase in the motion
of the polymer chains which, in turn, produces a shift in
the hole-size distribution towards bigger holes. (Increased
energy enables the formation of larger gaps between adja-
Temperature
cent polymer chains due to loosened bonds.) By contrast,
Size selectivity within plant cuticles is very temperature the temporal abundance of small holes is relatively reduced
sensitive. For example, in ivy cuticles, the difference in at higher temperatures (Fig. 5).
solute mobilities between IAA (Vx=130 cm3 mol1) and An investigation of the response of solute mobility,
tebuconazole (Vx=241 cm3 mol1) amounts to factors of which is independent of time, revealed a large increase
24.5, 10.5, and 2.2 at 25 8C, 35 8C, and 55 8C, respectively with rising temperature (Baur and Schönherr, 1995;
(Buchholz, 1998). Therefore, size selectivity diminishes at Knoche and Bukovac, 2001; Schreiber, 2002). The tem-
elevated temperatures and intensifies at low temperatures. perature coefficients Q10 ranged from 3 (IAA in Prunus
Lipophilic pathway 2507
armeniaca) to 14 (cholesterol in Hedera helix) at 20 8C T (°C)
(Baur et al., 1997a). The temperature effect can be quan- 60 50 40 30 20
-3.5
tified using the Arrhenius equation:
Pyrus cv. ‚Gellerts B.‘ / Tebuconazole
k* = k*0 3expðED =RTÞ ð4Þ -4.0
1
where k*0 is the pre-exponential factor, ED (kJ mol ) is
the activation energy of diffusion, R (J mol1 K1) the -4.5
gas constant, and T (K) the temperature. By plotting logk*
(s1) versus 1/T (K1) the activation energy of diffusion

log k*
(ED) can be calculated from the slope (b) of the graph: -5.0

ED = 13b3R32:3 ð5Þ
-5.5

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


The Arrhenius graphs published were linear in the phy-
siologically relevant temperature range of 10–45 8C, show- sonication
-6.0
ing that ED was constant and that no distinct phase transitions check
caused by sudden structural changes occurred. At temper-
atures above 50 8C, a slight curvature of Arrhenius graphs -6.5
towards lower ED occurred for some species. This was an 3.0 3.1 3.2 3.3 3.4
additional hint that cuticles could be classified as elasto- 1/T.1000(K-1)
mers, since this flattening quite often occurs with rubbery Fig. 6. Arrhenius graphs obtained for tebuconazole in CM of Pyrus
polymers, as ED values reflect the onset of viscous flow (cv. Gellerts Butterbirne). One sample of CMs (n=10) was treated
(Schlotter and Furlan, 1992). Depending on species and with sonication for 5 min in a water bath 1 d before experimental use,
whereas the other sample (n=11) remained untreated.
solute size, ED ranged from 60 to 232 kJ mol1 (Baur et al.,
1997a). This demonstrated that plant cuticles are very
effective diffusion barriers. ivation energy (enthalpy) of diffusion (ED). A strict linear
The structural and functional integrity of cuticles can correlation was obtained for all of the CMs from 14 plant
be examined by measuring mobility after temperature species and organic compounds having molar volumes
cycling. The comparison of subsequent experiments after ranging from 130 to 349 cm3 mol1 and cuticle/water
annealing can identify hysteresis effects due to re- partition coefficients of 18–108 (Fig. 7). This provided
arrangement of the wax constituents. Such experiments evidence that all of these compounds diffused—irrespec-
were conducted at temperatures up to 70 8C (Baur et al., tive of their size and lipophilicity—along the same
1997a). Results were plant-species specific and ranged lipophilic diffusion path. Each of the solutes must have
from superimposed graphs (Prunus laurocerasus) to travelled through a microenvironment with identical phys-
the straightening of previously curved graphs (Citrus icochemical properties (Buchholz and Schönherr, 2000).
aurantium), and in one case showed 10-fold increased There are indications that water and lipophilic solutes
mobilities at 25 8C (Hedera helix). These different results can share the lipophilic path. This co-permeability was
were related to differences in wax composition causing also confirmed by the simultaneous measurement of
3
different melting behaviours and redistribution during H-labelled water and 14C-labelled organic compounds
annealing. across plant cuticles (Niederl et al., 1998). Water can,
The sonication of CMs increased solute mobility at however, take two parallel paths of diffusion: the lipo-
25 8C by a factor of 4 and reduced the ED from123 kJ mol1 philic path made of cutin and wax and the polar path com-
to 92 kJ mol1. Phase transitions or structural defects posed of humidity-sensitive pores traversing the cuticles
were not indicated by the Arrhenius graphs (Fig. 6). The (Schreiber, 2002). Since wax extraction enhanced water
interpretation of these results remains speculative (e.g. permeability by several orders of magnitude (Schönherr,
loosening of chemical bonds), although a further observa- 1982), whereas the effect of humidity on cuticular water
tion was that these effects diminished during storage permeability was only 2–3-fold, it is reasonable to
(several months), and mobilities reverted to those of untreated argue that the largest fraction of water diffused across
CMs. This indicated a re-equilibration between the wax the lipophilic paths in the cuticle (Schreiber et al., 2001;
crystallites (‘healing of defects’) over time, as previously de- see also Table 2). The additional path predominated at
scribed for water transpiration (Geyer and Schönherr, 1990). temperatures higher than 30 8C and is likely to exhibit
Based on the transition state theory (Glasstone et al., a threshold in size and/or lipophilicity (Schreiber, 2002).
1941) data can be plotted according to the thermodynamic Arrhenius graphs obtained by repeat measurements
relationship between the pre-exponential factor (k*0) of on the same CM showed that, although the data points
the Arrhenius equation (proportional to entropy) and act- remained on the general line for CMs, they were shifted
2508 Buchholz
35
Citrus aurantium
Model Compounds:
IAA Citrus grandis
NAA
30 Bifenox Hedera helix
Tebuconazole
Cholesterol Ilex aquifolium

Ilex paraguariensis
25
log k*o Malus cv. Golden Delicious

Populus canescens
20
Prunus laurocerasus

Pyrus cv. Conference


15 Pyrus cv. Gellerts Butterbirne

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


Pyrus cv. Williams Christ
log k *o = -3.61 -1.28 ED 10-3 / R
10 Pyrus pyrifolia
r² = 0.99
Schefflera actinophylla

Strophanthus gratus
5
5 10 15 20 25 30 35
ED 10-3 / R

Fig. 7. Correlation between the Arrhenius slopes (ED/R) and y-intercepts of Arrhenius graphs k*0. Data for individual CMs from 14 plant species were
plotted. The solute mobilities were calculated for heating experiments with a temperature range of 25–50/55 8C for five different model compounds
(modified from Buchholz and Schönherr, 2000, with kind permission of Springer Science and Business Media).

towards lower entropy and enthalpy values (Fig. 8). A Baur et al., 1997b; Schönherr et al., 2001; Shi et al.,
consecutive third run rarely caused any further change in 2005), and also for active ingredients (Baur et al., 1996a).
this free energy relationship. Extraction of the cuticular This increased diffusivity is dependent on concentra-
waxes from the CMs greatly reduced the activation en- tion and is interpreted as an unspecific plasticizing effect
ergies and increased the solute mobilities. In the thermo- of the surfactant molecule on the ultrastructure of the wax
dynamic plot the linear dependency was maintained, which (Schreiber et al., 1996b). Rapid induction and complete
means that the free energy was identical in both membrane reversibility of the effects indicate that the physical wax
types, CM and MX. Additionally, the graphs obtained structure is not irreversibly altered (e.g. solubilization of
for MX exhibited a parallel displacement towards higher wax or destruction of crystalline wax domains) by the
entropy. This displacement was interpreted as a temperature- surfactant-induced acceleration (Schreiber, 1995, 2006).
independent tortuosity factor directly related to entropy The plasticizing effect of accelerators on cuticular
(Buchholz and Schönherr, 2000). Consecutive heating waxes in situ (CM) and on the cutin matrix (MX) was
cycles caused a shift along the line of a given membrane investigated by applying the entropy–enthalpy relationship
type towards lower entropy and enthalpy. This means that (Buchholz and Schönherr, 2000). The plasticizer tributyl
the individual chemical components might have been phosphate (TBP) increased solute mobilities and reduced
redistributed during the annealing phase but this did not ED significantly for both CM and MX. This indicated
affect the physicochemical properties of the diffusion path. that sorption of TBP increased segmental mobility of
polyethylene chains in both cutin and wax. This effect
was almost completely reversible, and for both plasticized
Plasticizer effects
membrane types (CM, MX) the thermodynamic relation-
Foliar uptake can be affected by formulation and adjuv- ship existed whereby all data points lay on the same line.
ants (Falk, 1994; Baur, 1998). Formulants are any added These data were used to characterize the lipophilic path-
material in a pesticide formulation other than the biol- way across plant cuticles in terms of the free-volume
ogically active ingredient(s), whereas adjuvants are those theory, giving evidence that both high temperature and
formulants designed to enhance the activity or other accelerators act similarly on the barrier properties of a CM.
properties of an agrichemical product (Holland, 1996). It is known that the efficacy of an accelerator is
Surfactants that enhance foliar uptake by increasing dependent on both the type of plasticizer and on its concen-
the mobility of a given compound are called accelerating tration. Arrhenius diagrams indicate that lower activation
adjuvants (Schönherr, 1993b). This accelerator effect has energies are produced with higher loadings of plasticizers
been demonstrated for various surfactants which were also (Schönherr et al., 2001). The different intrinsic efficacies
plasticizers (Schönherr, 1993a; Schönherr and Baur, 1996; of two plasticizers have been elucidated using this method.
Lipophilic pathway 2509
20 residue). If these passive adjuvants are hygroscopic, their
Strophanthus gratus / IAA effect on driving forces depends greatly on humidity.
Active adjuvants are solvents as well, but in addition they
penetrate into cuticles and increase solute mobility. These
15 effects depend on the type of accelerator, temperature,
size of solute, and plant species. Accelerators decrease the
viscosity of amorphous waxes which results in reduced
size selectivity. Accelerators also reduce differences in
log k*o

10
solute mobility among plant species (see above).

MX 1st run Foliar uptake


MX 2nd run
5 When foliar uptake of lipophilic solutes is considered,

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


CM 1st run
one has to be aware of the three consecutive steps: sorption,
CM 2nd run diffusion, desorption (see above). Generally, the rates of
CM 3rd run cuticular penetration (J) can be modelled in terms of
0 solute mobilities in the limiting skin and driving forces
0 5 10 15 20 (Schönherr and Baur, 1994; Schönherr et al., 1999):
ED 10-3 / R
J = M=ðAtÞ = k*3lls 3ðKwxfr Cfr  Kmxw Cw Þ ð6Þ
Fig. 8. Correlation between Arrhenius slopes (ED/R) and y-intercepts
of Arrhenius graphs k*0. Data for individual cuticular membranes (CM) The amount (M) of active substance penetrating per area
and polymer matrix membranes (MX) from Strophanthus gratus leaves
are plotted. The double arrow indicates the parallel displacement of (A and time (t) is the rate of penetration (J). The term in
both lines with identical slope. This displacement amounted to 1.7 brackets denotes the driving force across the CM.
(log scale). Solute mobilities were calculated for two or three succes- The sorption process is especially affected by numerous
sive heating experiments with identical membranes at temperature
ranges of 25–55 8C (CM) and 15–30 8C (MX). Between two tem- factors which change during droplet drying (or rewetting)
perature scans the membranes were kept at ambient temperature for 24 h. and can rarely be analysed individually. Figure 9 attempts
to illustrate the most relevant factors. Physicochemical
properties of the active substance (group 1) affect the
‘external’ term of the driving force: KwxfrCfr. For example,
This also demonstrated that the penetrating compound low water solubility causes a low concentration of dis-
and the surfactant must be present simultaneously in the solved molecules (Cfr), and will result in minor sorption
limiting skin in order to enhance the compound’s mobility. in cuticular wax (Kwxfr), although partitioning could
Perkins et al. (2005) demonstrated the plasticizing ef- be high. But a spray solution comprises, beside the active
fect on both reconstituted waxes and on the native leaf sur- substance, adjuvants and other formulants, like emulsifiers,
face. Localized thermal analysis measurements revealed buffer, etc. (group 2) which affect, for example, solvation
considerable variation in the wax melting point (Tm) over and partitioning (Baur et al., 1997c). Further, environ-
the leaf surface, indicating lateral heterogeneity in wax mental factors like temperature and humidity govern the
composition. The authors proposed areas of amorphous velocity of these dynamic processes. Consequently, it has
wax formations of at least 1 lm2. After droplet application to be realized that, in order to maximize rates of pen-
of Synperonic A7 (ethoxylated alcohol) they observed a etration, the product of Kwxfr and Cfr has to be increased,
depression in the wax melting point, mainly at the deposit whereas both quantities are mutually dependent on each other.
annulus, whereas the surfactant Synperonic A20 (with a The individual parameters affecting solute mobility
more homogenous deposit structure) had no observable (group 3) were extensively described above. After travel-
effect on the Tm of the wax. ling through the cuticle the ‘internal’ term of the driving
According to the model presented to describe the foliar force, KmxwCw becomes relevant. Here, the partitioning
uptake of non-electrolytes (Schönherr et al., 1999), these in the aqueous phase of the apoplast and potential access
two surfactants could be classified as active (accelerator; to the symplast (after traversing the biomembrane) de-
plasticizing) or passive (donor; dissolving) adjuvants termine the translocation into plant tissues. Rapid distri-
(Cronfeld et al., 2001). Passive adjuvants, which remain bution will result in a low concentration underneath the
on the surface of the cuticles, can serve as solvents, and spray droplet and enable large driving forces.
they affect driving forces via their effects on Kwxfr Finally, all those parameters describing the lipophilic
(partitioning of the active substance or adjuvants between pathway of non-electrolytes across plant cuticles—and
the formulation residue and the cuticular waxes) and Cfr demarcating it from the aqueous pathway (Schönherr,
(concentration of the active substance in the formulation 2006)—are summarized in Table 2.
2510 Buchholz
Conclusions 2006) allowed the quantification and interpretation of
size selectivity and tortuosity, as well as entropy/enthalpy
It is the big advantage of the UDOS method that, with
relationships. Therefore, the properties of the lipophilic
this desorption measurement and the determination of
pathway are now completely characterized. Only the
the solute mobility k*, a pure descriptor of the diffusion
interpretation of the plant species-specific y-intercept of
process in CMs can be obtained. Only the feasibility of
a hypothetical molecule having zero molar volume, k*0,
determining sorption-independent mobilities in CMs
and its reference to tortuosity, remain to be explained
(or diffusion coefficients in extracted wax; Schreiber,
in more detail.

Table 2. Comparison of the lipophilic and the aqueous pathways across plant cuticles
Lipophilic pathway Aqueous pathway

Plant Epicuticular waxes (Retention on leaf surface)

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


No effect on permeability Limit access to aqueous pores
Intracuticular waxes Amorphous waxes determine viscosity Minor effect on penetration
Crystallites cause tortuosity and
reduction in free volume
Cutin matrix Cross-linkage affects mobility Quality and spatial arrangement of polar
constituents determine abundance of aqueous pores
Substance Molecular size Pronounced size selectivity Hydrate shells enforce weak size selectivity
(no plasticizer)
Lipophilicity Affects driving force (sorption in waxes) –
No effect on mobility
Point of deliquescence – Low POD reduces risk of crystallized
formulation residues
Concentration Effect on driving force Effect on driving force
Microclimate Temperature Effect on fluidity, viscosity of waxes Minor effect on penetration
and matrix
Humidity Effect on driving force (solubility) Effect on driving force (solubility) and
swelling of cuticle
Passive additives Surfactants (Retention; mass flow into open stomata)
Effect on driving force (solvation, partitioning) Facilitation of access to aqueous pores
Minor effect on penetration
Humectants Effect on driving force (partitioning, solvation, Effect on driving force (solubility,
concentration dilution) concentration dilution)
Prevention of crystallization extends time Prevention of crystallization facilitates
for penetration access to aqueous pores and extends time
for penetration
Active additives Plasticizer Penetrate into cuticle and decrease viscosity Negligible effect on penetration
(accelerator)
Concentration dependent efficacy
Plasticizer’s and substance’s mobility need
a mutual fit

Fig. 9. Schematic drawing of a spray droplet and the underlying plant epidermis. Factors affecting cuticular penetration via the ‘lipophilic pathway’ are
grouped into (1) physicochemical properties of the active substance, (2) conditions determined by the environment and the spray solution, and (3)
parameters affecting solute mobility, k*. J, Rate of cuticular penetration; lls, length of the diffusion path in the limiting skin (6¼ thickness); Kwxfr, partition
coefficient wax/formulation residue; Kmxw, partition coefficient polymer matrix/water; Cfr, concentration of active substance in formulation residue;
Cw, concentration of active substance in water (apoplast) (according to the model described by Schönherr et al., 1999).
Lipophilic pathway 2511
This knowledge on factors affecting solute mobility Chaiko DJ, Leyva AA. 2005. Thermal transition and barrier
(group 3 in Fig. 9) and the physicochemical properties of properties of olefinic nanocomposites. Chemical Materials 17,
13–19.
a given non-polar active substance (group 1 in Fig. 9) are
Cronfeld P, Lader K, Baur P. 2001. Classification of adjuvants
the starting points for tailored crop protection products. and adjuvant blends by effects on cuticular penetration. ASTM
Special Technical Publication STP 1400 (Pesticide Formulations
and Application Systems: 20th Volume), 81–94.
References Coughlin CS, Mauritz KA, Storey RF. 1991. A general free
volume based theory for the diffusion of large molecules in
Baker EA. 1982. Chemistry and morphology of plant epicuticular amorphous polymers above Tg. 4. Polymer-penetrant interactions.
waxes. In: Cutler DF, Alvin KL, Price CE, eds. The plant cuticle. Macromolecules 24, 1526–1534.
London: Academic Press, 139–166. Cussler EL, Hughes SE, Ward WJ, Rutherford A. 1988. Barrier
Baur P. 1998. Mechanistic aspects of foliar penetration of agro- membranes. Journal of Membrane Science 38, 161–174.
chemicals and the effects of adjuvants. Recent Research Develop- Duda JL, Zielinski JM. 1996. Free-volume theory. In: Neogi, ed.
ments in Agricultural and Food Chemistry 2, 809–837. Diffusion in polymers. New York, NY: Marcel Dekker, 143–171.
Baur P, Buchholz A, Schönherr J. 1997a. Diffusion in plant Falk RH. 1994. Influence of formulation and adjuvants on the

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


cuticles as affected by temperature and size of organic solutes: foliar location and physical form of the active ingredient. In:
similarity and diversity among species. Plant, Cell and Environ- Holloway PJ, Rees RT, Stock D, eds. Interaction between ad-
ment 20, 982–994. juvants, agrochemicals and target organisms. Berlin: Springer-
Baur P, Grayson BT, Schönherr J. 1996a. Concentration-dependent Verlag, 53–82.
mobility of chlorfenvinphos in isolated plant cuticles. Pesticide Geyer U, Schönherr J. 1990. The effect of the environment on
Science 47, 171–180. the permeability and composition of Citrus leaf cuticles. I.
Baur P, Grayson BT, Schönherr J. 1997b. Polydisperse ethoxy- Water permeability of isolated cuticular membranes. Planta 180,
lated fatty alcohol surfactants as accelerators of cuticular penetra- 147–153.
tion. 1. Effects of ethoxy chain length and the size of the Glasstone S, Laidler K, Eyring H. 1941. The theory of rate
penetrants. Pesticide Science 51, 131–152.
processes. New York, NY: McGraw-Hill.
Baur P, Marzouk H, Schönherr J, Bauer H. 1996b. Mobilities
Haas K, Schönherr J. 1979. Composition of soluble cuticular
of organic compounds in plant cuticles as affected by structure
lipids and water permeability of cuticular membranes from Citrus
and molar volumes of chemicals and plant species. Planta 199,
leaves. Planta 146, 399–403.
404–412.
Hauke V, Schreiber L. 1998. Ontogenetic and seasonal develop-
Baur P, Marzouk H, Schönherr J, Grayson BT. 1997c. Partition
ment of wax composition and cuticular transpiration of ivy
coefficients between plant cuticle and adjuvants as related to
(Hedera helix) sun and shade leaves. Planta 207, 67–75.
foliar uptake. Journal of Agricultural and Food Chemistry 45,
Hoad SP, Jeffree CE, Grace J. 1992. Effects of wind and abrasion
3659–3665.
Baur P, Marzouk H, Schönherr J. 1999. Estimation of path lengths on cuticular integrity in Fagus sylvatica L. and consequences
for diffusion of organic compounds through leaf cuticles. Plant, for transfer of pollutants through leaf surfaces. Agriculture
Cell and Environment 22, 291–299. Ecosystems and Environment 42, 275–289.
Bauer H, Schönherr J. 1992. Determination of mobilities of organic Holland PT. 1996. Glossary of terms relating to pesticides
compounds in plant cuticles and correlation with molar volumes. (IUPAC Recommendations 1996). Pure and Applied Chemistry
Pesticide Science 35, 1–11. 68, 1167–1193.
Baur P, Schönherr J. 1995. Temperature dependence of the dif- Holloway PJ. 1982. Structure and histochemistry of plant cuticular
fusion of organic compounds across plant cuticles. Chemosphere membranes: an overview. In: Cutler DF, Alvin KL, Price CE, eds.
30, 1331–1340. The plant cuticle. London: Academic Press, 45–85.
Becker M, Kerstiens G, Schönherr J. 1986. Water permeability Holloway PJ. 1993. Structure and chemistry of plant cuticles.
of plant cuticles: permeance, diffusion and partition coefficients. Pesticide Science 37, 203–206.
Trees 1, 54–60. Holloway PJ. 1994. Plant cuticles: physicochemical characteristics
Bianchi G. 1995. Plant waxes. In: Hamilton RJ, ed. Waxes: and biosynthesis. In: Percy KE, Cape JN, Jagels R, Simpson CJ,
chemistry, molecular biology and functions. Dundee: The Oily eds. Air pollutants and the leaf cuticle. NATO ASI Series, Series
Press, 175–222. G: Ecological Sciences 36, 1–13.
Buchholz A. 1998. Größenselektivität und Temperaturabhängigkeit Jeffree CE. 1986. The cuticle, epicuticular waxes and trichomes
der Diffusion organischer Verbindungen in der pflanzlichen of plants, with reference to their structure, functions and evolution.
Kutikula. PhD thesis, University of Kiel, Germany. In: Juniper BE, Southwood TRE, eds. Insects and the plant
Buchholz A, Baur P, Schönherr J. 1998. Differences among plant surface. London: Edward Arnold, 23–63.
species in cuticular permeabilities and solute mobilities are not Jeffree CE. 1996. Structure and ontogeny of plant cuticles. In:
caused by differential size selectivities. Planta 206, 322–328. Kerstiens G, ed. Plant cuticles – an integrated functional
Buchholz A, Schönherr J. 2000. Thermodynamic analysis of dif- approach. Oxford: BIOS Scientific Publishers, 33–75.
fusion of non-electrolytes across plant cuticles in the presence and Jetter R, Schäffer S, Riederer M. 2000. Leaf cuticular waxes
absence of the plasticiser tributyl phosphate. Planta 212, 103–111. are arranged in chemically and mechanically distinct layers: evid-
Bukovac MJ, Flore JA, Baker EA. 1979. Peach leaf surfaces: ence from Prunus laurocerasus L. Plant, Cell and Environment
changes in wettability, retention, cuticular permeability, and epi- 23, 619–628.
cuticular wax chemistry during expansion with special reference to Juniper BE. 1995. Waxes on plant surfaces and their interactions
spray application. Journal of the American Society for Horticul- with insects. In: Hamilton RJ, ed. Waxes: chemistry, molecular
tural Science 104, 611–617. biology and functions. Dundee: The Oily Press, 157–176.
Carreto A, Almeida AR, Fernandes AC, Vaz WLC. 2002. Kerstiens G. 1996. Signalling across the divide: a wider perspec-
Thermotropic mesotropism of a model system for the plant tive of cuticular structure: function relationships. Trends in Plant
epicuticular wax layer. Biophysical Journal 82, 530–540. Science 1, 125–129.
2512 Buchholz
Kirkwood RC. 1999. Recent developments in our understanding Riederer M, Schneider G. 1990. The effect of the environment
of the plant cuticle as a barrier to the foliar uptake of pesticides. on the permeability and composition of Citrus leaf cuticles. II.
Pesticide Science 55, 69–77. Composition of soluble cuticular lipids and correlation with
Kirsch T, Kaffarnik F, Riederer M, Schreiber L. 1997. Cuticular transport properties. Planta 180, 154–165.
permeability of the three tree species Prunus laurocerasus L., Ginkgo Riederer M, Schreiber L. 1995. Waxes: the transport barriers of
biloba L., and Juglans regia L.: comparative investigation of the plant cuticles. In: Hamilton RJ, ed. Waxes: chemistry, molecular
transport properties of intact leaves, isolated cuticles, and reconstituted biology and functions. Dundee: The Oily Press, 131–156.
cuticular waxes. Journal of Experimental Botany 48, 1035–1045. Riederer M, Schreiber L. 2001. Protecting against water loss:
Knoche M, Bukovac MJ. 2001. Finite dose diffusion studies. III. analysis of the barrier properties of plant cuticles. Journal of
Effects of temperature, humidity and deposit manipulation on Experimental Botany 52, 2023–2032.
NAA penetration through isolated tomato fruit cuticles. Pest Santier S, Chamel A. 1998. Reassessment of the role of cuticular
Management Science 57, 737–742. waxes in the transfer of organic molecules through plant cuticles.
Knoche M, Petracek PD, Bukovac MJ. 2000. Finite dose diffu- Plant Physiology and Biochemistry 36, 225–231.
sion studies. I. Characterizing cuticular penetration in a model Schlotter NE, Furlan PY. 1992. A review of small molecule
system using NAA and isolated tomato fruit cuticles. Pest Man- diffusion in polyolefins. Polymers 33, 3323–3342.
agement Science 56, 1005–1015. Schmidt HW, Merida T, Schönherr J. 1981. Water permeability

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


Kolattukudy PE. 1984. Biochemistry and function of cutin and and fine structure of cuticular membranes isolated enzymatically
suberin. Canadian Journal of Botany 62, 2918–2933. from leaves of Clivia miniata Reg. Zeitschrift für Pflanzen-
Kolattukudy PE, Rogers LM, Li D, Hwang C, Flaishman MA. physiologie 105, 41–51.
1995. Surface signaling in pathogenesis. Proceeding of the Schönherr J. 1982. Resistance of plant surfaces to water loss:
National Academy of Sciences, USA 92, 4080–4087. transport properties of cutin, suberin and associated lipids. In:
Krüger H, van Rensburg L, Peacock J. 1996. Cuticular membrane Lange OL, Nobel PS, Osmond CB, Ziegler H, eds. Encyclopaedia
fine structure of Nicotiana tabacum L. leaves. Annals of Botany of plant physiology, New Series, Vol. 12B. Berlin: Springer,
77, 11–16. 153–179.
Leece DR. 1976. Composition and ultrastructure of leaf cuticles Schönherr J. 1993a. Effects of monodisperse alcohol ethoxylates
from fruit trees, relative to differential foliar absorption. on mobility of 2,4-D in isolated plant cuticles. Pesticide Science
Australian Journal of Plant Physiology 3, 833–847. 38, 155–164.
Lieb WR, Stein WD. 1972. The molecular basis of simple diffusion Schönherr J. 1993b. Effects of alcohols, glycols and monodisperse
within biological membranes. In: Bonner F, Kleinzeller A, eds. ethoxylated alcohols on mobility of 2,4-D in isolated plant cuticles.
Current topics in membrane transport, Vol. 2. New York, NY: Pesticide Science 39, 213–223.
Academic Press, 1–39. Schönherr J. 2006. Characterization of aqueous pores in plant
Marga F, Pesacreta TC, Hasenstein KH. 2001. Biochemcial cuticles and permeation of ionic solutes. Journal of Experimental
analysis of elastic and rigid cuticles of Cirsium horridulum. Planta Botany 57, 2471–2491.
213, 841–848. Schönherr J, Baur P. 1994. Modeling penetration of plant cuticles
McGowan JC. 1969. The effects of pressure and temperature on by crop protection agents and effects of adjuvants on their rates
the densities of liquid polymers. Polymer – the Science and Tech- of penetration. Pesticide Science 42, 185–208.
nology of Polymers 10, 841–848. Schönherr J, Baur P. 1996. Effects of temperature, surfactants and
McGowan JC, Mellors A. 1986. Molecular volumes in chemistry other adjuvants on rates of uptake of organic compounds. In:
and biology: applications including partitioning and toxicity. Kerstiens G, ed. Plant cuticles: an integrated functional
Chichester: Ellis Horwood. approach. Oxford: BIOS Scientific Publishers, 134–154.
Merk S, Blume A, Riederer M. 1998. Phase behavior and Schönherr J, Baur P, Buchholz A. 1999. Modelling foliar
crystallinity of plant cuticular waxes studied by Fourier transform penetration: its role in optimizing pesticide delivery. In: Brooks
infrared spectroscopy. Planta 204, 44–53. GT, Roberts TR, eds. Pesticide chemistry and bioscience, the
Niederl S, Kirsch T, Riederer M, Schreiber L. 1998. Co- food-environment challenge. Cambridge: The Royal Society of
permeability of 3H-labeled water and 14C-labeled organic acids Chemistry, 134–151.
across isolated plant cuticles: investigating cuticular paths of dif- Schönherr J, Riederer M. 1988. Desorption of chemicals from
fusion and predicting cuticular transpiration. Plant Physiology plant cuticles: evidence for asymmetry. Archives of Environmental
116, 117–123. Contamination and Toxicology 17, 13–19.
Orgell WH. 1955. The isolation of plant cuticle with pectic enzymes. Schönherr J, Schmidt HW. 1979. Water permeability of plant
Plant Physiology 30, 78–80. cuticles: dependence of permeability coefficients of cuticular
Perkins MC, Roberts CJ, Briggs D, Davies MC, Friedmann A, transpiration on vapor pressure saturation deficit. Planta 144,
Hart C, Bell G. 2005. Macro- and microthermal analysis of plant 391–400.
wax/surfactant interactions: plasticizing effects of two alcohol Schönherr J, Schreiber L, Buchholz A. 2001. Effects of temper-
ethoxylated surfactants on an isolated cuticular wax and leaf ature and concentration of the accelerators ethoxylated alcohols,
model. Applied Surface Science 243, 158–165. diethyl suberate and tributyl phosphate on the mobility of [14C]2,
Petracek PD, Bukovac MJ. 1995. Rheological properties of 4-dichlorophenoxy butyric acid in plant cuticles. Pest Manage-
enzymatically isolated tomato fruit cuticle. Plant Physiology ment Science 57, 17–24.
109, 675–679. Schreiber L. 1995. A mechanistic approach towards surfactant/
Reynhardt EC, Riederer M. 1991. Structure and molecular wax interactions: effects of octaethyleneglycolmonododecylether
dynamics of the cuticular wax from leaves of Citrus aurantium on sorption and diffusion of organic chemicals in reconstituted
L. Journal of Physics D: Applied Physics 24, 478–486. cuticular wax of barley leaves. Pesticide Science 45, 1–11.
Reynhardt EC, Riederer M. 1994. Structures and molecular Schreiber L. 2002. Co-permeability of 3H-labelled water and
14
dynamics of plant waxes. II. Cuticular waxes from the leaves of C-labelled organic acids across isolated Prunus laurocerasus
Fagus sylvatica L. and Hordeum vulgare L. European Biophysics cuticles: effect of temperature on cuticular paths of diffusion.
Journal 23, 59–70. Plant, Cell and Environment 25, 1087–1094.
Lipophilic pathway 2513
Schreiber L. 2006. Sorption and diffusion of lipophilic molecules Shi T, Simanova E, Schönherr J, Schreiber L. 2005. Effects
in cuticular wax and effects of accelerators on solute mobilities. of accelerators on mobility of 14C-2,4-dichlorophenoxy butyric
Journal of Experimental Botany 57, 2515–2523. acid in plant cuticles depends on type and concentration of
Schreiber L, Kirsch T, Riederer M. 1996a. Transport properties of accelerator. Journal of Agricultural and Food Chemistry 53,
cuticular waxes of Fagus sylvatica L. and Picea abies (L.) Karst.: 2207–2212.
estimation of size selectivity and tortuosity from diffusion Tamura H, Knoche M, Bukovac MJ. 2001. Evidence for surfactant
coefficients of aliphatic molecules. Planta 198, 104–109. solubilization of plant epicuticular wax. Journal of Agricultural
Schreiber L, Krimm U, Knoll D. 2004. Interactions between and Food Chemistry 49, 1809–1816.
epiphyllic microorganisms and leaf cuticles. In: Varma A, Abbott Turunen M, Huttunen S. 1990. A review of the response of
L, Werner D, Hampp R, eds. Plant surface microbiology, 145–156. epicuticular wax of conifer needles to air pollution. Journal of
Schreiber L, Riederer M. 1996. Determination of diffusion Environmental Quality 19, 35–45.
coefficients of octadecanoic acid in isolated cuticular waxes and Viougeas MA, Rohr R, Chamel A. 1995. Structural changes and
their relationship to cuticular water permeabilities. Plant, Cell permeability of ivy (Hedera helix L.) leaf cuticles in relation to
and Environment 19, 1075–1082. leaf development and after selective chemical treatments. New
Schreiber L, Riederer M, Schorn K. 1996b. Mobilities of organic Phytologist 130, 337–348.
von Wettstein-Knowles P. 1995. Biosynthesis and genetics of

Downloaded from https://academic.oup.com/jxb/article/57/11/2501/677149 by guest on 31 March 2021


compounds in reconstituted cuticular wax of barley leaves: effects
waxes. In: Hamilton RJ, ed. Waxes: chemistry, molecular biology
of monodisperse alcohol ethoxylates on diffusion of pentachloro-
and functions. Dundee: The Oily Press, 91–130.
phenol and tetracosanoic acid. Pesticide Science 48, 117–124. Watanabe A. 1961. Synthesis and physical properties of normal
Schreiber L, Schönherr J. 1993. Mobilities of organic compounds higher primary alcohols. IV. Thermal and X-ray studies on the
in recrystallized cuticular wax of barley leaves: determination of polymorphism of the alcohols of even carbon numbers from
diffusion coefficients. Pesticide Science 38, 353–361. dodecanol to tetratriacontanol. Bulletin of the Chemical Society
Schreiber L, Schorn K, Heimburg T. 1997. 2H NMR study of of Japan 34, 1728–1734.
cuticular wax isolated from Hordeum vulgare L. leaves: identifi- Wattendorf J, Holloway PJ. 1980. Studies on the ultrastructure
cation of amorphous and crystalline wax phases. European Bio- and histochemistry of plant cuticles: the cuticular membrane of
physics Journal 26, 371–380. Agave americana L. in situ. Annals of Botany 46, 13–28.
Schreiber L, Skrabs M, Hartmann K, Diamantopoulos P, Whitehouse P, Holloway PJ, Caseley JC. 1982. The epicuticular
Simanova E, Santrucek J. 2001. Effect of humidity on cuticular wax of wild oats in relation to foliar entry of the herbicides
transpiration of isolated cuticular membranes and leaf disks. diclofop-methyl and difenzoquat. In: Cutler DF, Alvin KL, Price
Planta 214, 274–282. CE, eds. The plant cuticle. Linnean Society Symposium Series 10.
Shafer WE, Schönherr J. 1985. Accumulation and transport of London: Academic Press, 315–330.
phenol, 2-nitrophenol, and 4-nitrophenol in plant cuticles. Ecotox- Wiedemann P, Neinhuis C. 1998. Biomechanics of isolated plant
icology and Environmental Safety 10, 239–252. cuticles. Botanica Acta 111, 28–34.

You might also like