Running Head: Label-free in vivo SRS analysis of plant cuticle

advertisement
1
Running Head: Label-free in vivo SRS analysis of plant cuticle.
2
3
Corresponding author (Stimulated Raman Scattering Microscopy and Submission):
4
Julian Moger, Physics and Astronomy, Physics Building, College of Engineering, Mathematics
5
and Physical Sciences, University of Exeter, Stocker Road, Exeter, U.K.
6
EX4 4QL.
7
+44 (0) 1392 724181
8
J.Moger@exeter.ac.uk
9
10
Corresponding author (Plant Cuticle): John Love, Biosciences, Geoffrey Pope Building,
11
College of Life and Environmental Sciences, University of Exeter, Stocker Road, Exeter, U.K.
12
EX4 4QD.
13
+44 (0) 1392 725169
14
J.Love@exeter.ac.uk
15
16
17
Research Area: Breakthrough Technologies.
18
1
19
IN VIVO CHEMICAL AND STRUCTURAL ANALYSIS OF PLANT CUTICULAR WAXES USING STIMULATED
20
RAMAN SCATTERING (SRS) MICROSCOPY
21
22
George R. Littlejohn1*, Jessica C. Mansfield2*, David Parker3, Rob Lind4, Sarah Perfect4, Mark
23
Seymour4, Nicholas Smirnoff1, John Love1 and Julian Moger2
24
25
1
26
Exeter, Stocker Road, Exeter, Devon, EX4 4QD. U.K.
Biosciences, Geoffrey Pope Building, College of Life and Environmental Sciences, University of
27
28
2
29
Sciences, Stocker Road, University of Exeter, Exeter, Devon, EX4 4QL. U.K.
Physics and Astronomy, Physics Building, College of Engineering, Mathematics and Physical
30
31
3
32
Westhollow Technology Center, 3333 Highway 6 South, Houston, TX 77082-3101. U.S.A.
Biodomain
Technology Group,
Shell
International
Exploration
and
Production
Inc.,
33
34
Syngenta, Jealott’s Hill International Research Centre, Bracknell, Berkshire, RG42 6EY U.K.
4
35
36
*These authors contributed equally to the research.
37
38
One-Sentence Summary of the Research: Stimulated Raman Microscopy is a new in vivo
39
imaging technique that enables simultaneous chemical and structural analysis of plant cuticle.
40
41
Footnotes: This work was supported by grants from the Biotechnology and Biological Sciences
42
Research Council (J.L, J.M), Shell Research Ltd. (J.L), and Syngenta (J.M).
43
2
44
Abstract
45
The cuticle is a ubiquitous, predominantly waxy layer on the aerial parts of higher plants that
46
fulfils a number of essential physiological roles, including regulating evapotranspiration, light
47
reflection and heat tolerance, control of development, and providing an essential barrier
48
between the organism and environmental agents such as chemicals or some pathogens. The
49
structure and composition of the cuticle are closely associated, but are typically investigated
50
separately using a combination of structural imaging and biochemical analysis of extracted
51
waxes.
52
including Fourier transform infrared spectroscopy (FTIR) microscopy and coherent anti-Stokes
53
Raman spectroscopy (CARS) microscopy, have been used to investigate the cuticle, but the
54
detection sensitivity is severely limited by the background signals from plant pigments.
55
We present a new method for label-free, in vivo structural and biochemical analysis of plant
56
cuticles based on stimulated Raman scattering (SRS) microscopy. As a proof-of-principle, we
57
used SRS microscopy to analyse the cuticles from a variety of plants, at different times in
58
development. We demonstrate the SRS virtually eliminates the background interference
59
compared to CARS imaging, and results in label-free chemically specific, confocal images of
60
cuticle architecture with simultaneous characterisation of cuticle composition. This innovative
61
use of the SRS spectroscopy may find applications in agrochemical R&D or in studies of wax
62
deposition during leaf development and, as such, represents an important step in the study of
63
the higher plant cuticle.
Recently, techniques that combine stain-free imaging and biochemical analysis,
64
3
65
Introduction
66
67
The Plant Cuticle
68
The majority of land plants possess an extracellular, waxy cuticle that covers the surface of their
69
aerial parts and protects them against desiccation, external physical and chemical stresses and
70
a variety of biological agents (Grncarevic and Radler 1967; Ristic and Jenks 2002; Krauss et al.
71
1997; Barthlott and Neinhuis, 1997; Yeats and Rose, 2013). The cuticle is a composite layer
72
composed mainly of cutin and overlaid by cuticular waxes. Cutin is a macromolecular structure
73
consisting primarily of hexadecanoic (palmitic) and octadecenoic (vaccenic) acids that are
74
covalently linked by ester bonds, to generate a rigid, three-dimensional network that is
75
embedded with polysaccharides. Cuticular waxes are composed of long-chain (C20-C40)
76
aliphatic molecules derived from fatty acids (Samuels et al. 2008) and studies over the last
77
several decades have identified structural and regulatory constituents of the biosynthetic
78
pathways of cuticular components (Kolattukudy 1981; Beisson et al. 2012). In addition to the
79
physio-chemical properties conferred by its lipid components, the architecture of the cuticle
80
plays an essential role in physiological function.
81
properties of the cuticular structure, the extraordinary super-hydrophobicity of the Lotus leaf has
82
been mimicked in micro and nano-technology to generate self-cleaning surfaces (Jung and
83
Bhushan 2006; Koch et al., 2009; Bhushan et al. 2009).
84
As may be expected given the diversity of plants, the habitats they inhabit and individual life-
85
histories, the morphology and composition of plant cuticle varies extensively between and within
86
species and includes plate, needle and pillar shaped wax crystals (Barthlott et al., 2008). In
87
some species, cuticular wax composition is known to vary with depth giving rise to chemically
88
distinguishable layers (Yeats and Rose, 2013). Finally, the cuticle is increasingly shown to be
89
important in development (Koornneef et al., 1989; Yeats and Rose, 2013) and pathogenesis
90
(Lee and Dean, 1994; Gilbert et al. 1996; Bessire et al., 2007; Delventhal et al., 2014). It is
91
therefore unsurprising that interest in cuticle composition, structure and physiology is increasing
92
(Heredia-Guerrero et al., 2014; Buschhaus et al., 2014; Xu et al., 2014; Hen-Aviv et al., 2014).
93
Moreover, a greater understanding of the relationship between structure and chemical
94
composition of cuticle waxes is vital for enhancing agriculture yields as it will further our
95
knowledge of how plants regulate water balance, and inform the application of nutrition (foliar
96
feeds) and pesticides, leading to improved formulation strategies for agrochemicals.
For example, through understanding the
97
98
The chemical composition and topological architecture of cuticular waxes are both critical for
99
optimal physiological function. Analyses of these essential properties have typically been
100
performed separately.
101
chromatography; cuticle ultrastructure may be analysed using destructive imaging techniques
Cuticle wax composition is normally determined using gas-
4
102
such as scanning electron microscopy (Baker and Holloway 1971; Jetter et al., 2000; Barthlott
103
et al. 2008) and laser desorption ionising mass spectroscopy (Jun et al., 2010) or, in vivo, using
104
non-destructive real-time techniques including white light scanning interferometry (Kim et al.,
105
2011), atomic force microscopy (Koch et al., 2004), confocal microscopy in reflectance mode
106
(Veraverbeke et al., 2001), fluorescence microscopy of chemical stains (Pighin, et al., 2004),
107
CARS microscopy (Yu et al., 2008; Weissflog et al., 2010) and total internal reflection Raman
108
spectroscopy (Greene and Bain, 2005). Despite the advances in our understanding of the
109
cuticle that have been made with these techniques, there is a great need for techniques that
110
combine chemical and structural information to provide in-situ high-resolution chemical analysis
111
of epicuticle waxes.
112
113
Techniques based on vibrational spectroscopy offer in-situ chemical analysis derived from the
114
vibrational frequencies of molecular bonds within a sample. However, due to water absorption
115
and the intrinsically low spatial resolution associated with the long infrared (IR) wavelengths
116
required to directly excite molecular vibrations, IR absorption techniques have limited value for
117
bio-imaging. Raman scattering however, provides analysis of vibrational frequencies by
118
examining the inelastic scattering of visible light. Raman scattered light is frequency-shifted with
119
respect to the incident light by discrete amounts that correspond to the vibrational frequencies
120
of molecular bonds within the sample. The spectrum of Raman scattered light consists of a
121
series of discrete peaks that each correspond to a molecular bond and can be regarded as a
122
chemical fingerprint holding a wealth of information regarding chemical composition.
123
Unfortunately, Raman scattering is an extremely weak effect and typical signals from biological
124
samples are at least six orders of magnitude weaker than those from fluorescent labels. This
125
severely limits the application of Raman for studying living systems since long acquisition times
126
(100 ms to 1 s per pixel) and relatively high excitation powers (several hundred mW) are
127
required to image most biomolecules with sufficient sensitivity. Furthermore, the lack of
128
sensitivity is compounded by autofluorescence, which in plant tissues completely overwhelms
129
the Raman signal, prohibiting its application in-planta.
130
131
Far stronger Raman signals can be obtained using coherent Raman scattering (CRS) (Min et al,
132
2011). CRS achieves a Raman signal enhancement by focusing the excitation energy onto a
133
specific molecular vibrational frequency (see Fig 1 (a)). A pump and Stokes beam, with
134
frequencies ωp and ωS respectively, are incident upon the sample with their frequency
135
difference (ωp – ωS) tuned to match the molecular vibrational frequency of interest (Ω). Under
136
this resonant condition the excitation fields efficiently drive bonds to produce a strong non-linear
137
coherent Raman signal. When applied in microscopy format the non-linear nature of the CRS
138
process confines the signal to a sub-micron focus that can be scanned in space, allowing three-
5
139
dimensional ‘confocal-like’ mapping of biomolecules. CRS microscopy has particular
140
advantages for bio-imaging:
141
(1) Chemically specific contrast is derived from the vibrational signature of endogenous
142
biomolecules with in the sample, negating the need for extraneous labels/stains.
143
(2) Low energy near-IR excitation wavelengths can be employed which reduce photodamage
144
and increase depth penetration into scattering tissues.
145
(3) The CRS process does not leave sample molecules in an excited state it does not suffer
146
from photobleaching and can be used for time-course studies.
147
148
Coherent Raman scattering microscopy may be achieved by detecting either coherent anti-
149
Stokes Raman scattering (CARS) or stimulated Raman scattering (SRS).
150
151
Coherent anti-Stokes Raman scattering (CARS) microscopy
152
CARS microscopy relies on detection of the anti-Stokes signal generated at frequency ωas =
153
2ωp – ωS, which, by using filters, is spectrally isolated from the pump and Stokes beams, and its
154
intensity used to map the location of biomolecules of interest (Zumbusch et al, 1999). The
155
CARS signal is blue-shifted with respect to the pump and Stokes wavelengths (see Fig. 1(b))
156
making CARS more resilient to sample autofluorescence than spontaneous Raman. However,
157
in highly autofluorescent samples such as plant tissues, the usually weak two-photon
158
fluorescence (also blue-shifted with respect to the excitation wavelengths) overwhelms the
159
CARS signal. Consequently CARS imaging in planta has only been applied to samples with
160
reduced chlorophyll autofluorescence including dried tissues (Zeng et al., 2010, Ding et al.,
161
2012), roots (Ly et al., 2007) and cuticle waxes after they have been stripped away from the leaf
162
(Weissflog et al., 2010).
163
164
Stimulated Raman Scattering (SRS) microscopy
165
SRS relies on detecting subtle changes in the intensities of the excitation fields that occur by
166
virtue of stimulated excitation (Freudiger et al, 2008). When the difference frequency, ωp – ωS,
167
matches the frequency of a molecular vibration the Stokes beam intensity (IS) experiences a
168
gain (ΔIS) while the intensity of the pump beam (Ip) experiences a loss (ΔIp), shown in Fig. 1 (c).
169
This transfer of intensity between the excitation beams only occurs when both beams are
170
incident simultaneously on the sample and is measured by detecting the modulation that is
171
transferred to the pump beam after it has passed through the sample when the intensity of the
172
Stokes beam is modulated. The amplitude of the transferred intensity modulation is directly
173
proportional to the concentration of target molecules and by modulating at frequencies above
174
laser noise (>1 MHz) can be detected with a lock-in amplifier with great sensitivity (Ye et al,
175
2009).
6
176
Since SRS is detected at the same wavelength as the excitation fields it is not affected by
177
fluorescent emission and as recently demonstrated by Mansfield (Mansfield et al., 2013), may
178
also be used in the presence of highly pigmented samples such as plant tissues.
179
180
We have recently shown that it is possible use phase-sensitive detection to separate the
181
vibrational SRS signal from the electronic absorption processes (Mansfield et al., 2013). In the
182
current investigation, we used SRS imaging to investigate both the structure and chemical
183
composition of plant cuticular waxes, simultaneously and in-vivo. We compared SRS images of
184
cuticle wax structures from variety of plant species that display characteristic wax morphologies
185
to SEM and CARS images of the leaf surface. We demonstrated that spectroscopic information
186
can be obtained to provide chemically-specific, in-situ analysis of waxes on living leaves with
187
submicron spatial resolution. Using SRS microscopy we compared the cuticle structure and
188
composition between wild type Arabidopsis thaliana and an eciferum-1 (cer1) mutant with
189
impaired cuticle deposition and have also shown that SRS imaging is sufficiently sensitive to
190
track temperature-induced changes in cuticle formation in the salt-cress, Thellungellia
191
salsuginea. We conclude that, with the increasing availability of commercial Raman-based
192
microscopes, SRS imaging has the potential to radically advance our understanding of plant
193
cuticular structure and composition, and their effects on physiology and development.
7
194
Results
195
196
Characterisation of the Raman Spectra of Dissolved Cuticle Waxes.
197
Spontaneous Raman spectra of hexane-extracted wax samples over the CH vibrational region
198
from 2700-3200cm-1 were acquired as a baseline reference for each plant species (Fig. 2).
199
Gaussian peaks were fitted underneath each of the spectra to visualise the relative contribution
200
of different molecular vibrations, based on peaks previously identified in the literature (Greene
201
and Bain, 2005, Snyder et al., 1978, Ho and Pemberton, 1998). The wax extracts from
202
Arabidopsis thaliana, Thellungiella parvula (Saltwater Cress), Musa acuminata (Banana) and
203
Monstera delicosa (Cheeseplant) had similar spectral profiles that matched those reported for
204
other waxes and alkanes reported in the literature (Snyder et al., 1978, Greene and Bain, 2005)
205
and deconvolved into 7 identifiable peaks (Table 1). The peaks, corresponding to the
206
symmetrical and anti-symmetrical CH2 stretch, are prominent on all spectra. The ratio anti-
207
symmetric to symmetric peak height is known to give a strong indication of the alkyl chain
208
conformational order with a ratio 1.6-2 indicating a highly crystalline structure and a ratio of 0.6-
209
0.9 indicating a liquid structure (Ho and Pemberton, 1998, Greene and Bain, 2005, Snyder et
210
al., 1978). For the wax samples analysed here, the ratios were as follows: A. thaliana 1.13, T.
211
parvula 1.58, M. acuminata 1.34 and M. delicosa 1.54, indicating structures intermediate
212
between a highly crystalline structure and a liquid. The curve fits to the spectra also included
213
two Fermi resonances of the CH2 bond one at 2930 cm-1 and another very broad Fermi
214
resonance at approximately 2870 cm-1 (Snyder et al., 1978). The contributions due to the CH3
215
symmetric and anti-symmetric stretches were small as proportionally there are much less of
216
these groups due to the long chain lengths. The fitting parameters for each wax along with the
217
goodness of fit are included in the supplementary information.
218
Two of the species investigated, Dudleya anthonyi (Dudleya) and Xerosicyos danguyi (Silver
219
Dollar Plant), showed dramatically different Raman spectra from A. thaliana, T. parvula, M.
220
acuminata and M. delicosa, characterised by additional peaks which are indicative of a more
221
complex chemical structure, with additional chemical bonds to those described in Table 1. In D.
222
anthonyi, the Raman spectrum changed depending on the number of the hexane wash and
223
hence, we surmise, on the depth of penetration into the cuticle; the initial, most superficial
224
hexane wash showed an unusual Raman spectrum, but subsequent washes that solubilise
225
waxes deeper in the cuticle showed spectra which matched more closely with those of a typical
226
alkane. These differences are most readily explained by both these species possessing a thick,
227
glaucous cuticle compared to the other mesophytes studied, and is an adaptation enabling
228
survival in xeric environments.
229
8
230
In-Vivo Comparison of Coherent anti-Stokes Raman scattering (CARS) and stimulated Raman
231
scattering (SRS) images of plant cuticle.
232
Having characterised the Raman spectra of waxes extracted from the cuticle of a variety of
233
plants, we imaged the cuticles of the same plants in vivo, using both coherent anti-Stokes
234
Raman scattering (CARS) microscopy (Weissflog et al, 2010) and stimulated Raman scattering
235
(SRS) microscopy (Fig. 3). SRS and CARS images of the epicuticular wax layer of T. parvula
236
leaves were acquired simultaneously (Fig. 3) and are presented as the exemplar to compare
237
the two techniques.
238
The
239
autofluorescence from cell walls and chloroplasts (Fig. 3A). This autofluorescence is due to
240
inevitable two photon-absorption by the compounds in these organelles, and the resulting
241
fluorescence emission at similar wavelengths to the CARS signal. Consequently, in the CARS
242
images, only the largest wax crystals can be visualized.
243
Conversely, in the SRS image (Fig. 3B) there is no fluorescent background because the signal
244
measured is not due to a change in wavelength (i.e. fluorescence relative to excitation), but a
245
change in intensity between the excitation lasers (the pump and Stokes beams) following
246
stimulated Raman excitation of the samples. The cuticular wax crystals are clearly visible and
247
the SRS signal spectrally matches that of the spontaneous Raman spectrum of purified cuticular
248
waxes (Fig. 3C). Cell walls are also visible in the SRS images, although with a lesser intensity
249
than the wax crystals, which can be explained by the overlap between the Raman spectra of the
250
cuticular waxes and cellulose. SRS microscopy therefore provides not only clearer images of
251
plant cuticle structure than CARS microscopy, but, due to the low-levels of background noise in
252
the images, also has the capacity to yield information on specific compounds that compose the
253
imaged structures.
images
acquired
using
CARS
are
dominated
by
high-levels
of
background
254
255
Comparison of stimulated Raman scattering (SRS) and Scanning Electron Microscope Images
256
of Plant Cuticle.
257
To confirm the capacity of SRS microscopy for accurately resolving cuticular structures, the
258
surface of A. thaliana stems were imaged using SRS microscopy and the more conventional
259
scanning electron microscopy (SEM; Fig. 4). SRS and SEM images of untreated, wild-type
260
stems (Fig. 4A and 4B) show an uneven surface composed of similarly-proportioned globules,
261
although the level of detail is higher in the SEM image. We ascertained that these globular
262
structures were indeed waxes by washing the stems in hexane prior to imaging (Fig. 4C and
263
4B). In this case, both images show that the surface of the stem is smooth, with clearly visible
264
cells and stomata. Finally, stems from A. thaliana defective for ECERIFERUM1 (CER1) gene
265
expression were imaged using either technique (Fig. 4E and 4F). cer1 mutants are impaired in
9
266
cuticle wax biosynthesis (Aarts et al. 1995; Bernard et al. 2012), and the images indeed show
267
stems with markedly reduced surface structures than the wild-type (Fig. 4A and 4B).
268
Unlike SEM which provides information only on the surface topology of imaged samples, SRS
269
microscopy is a technique of multi-photon confocal microscopy that, like conventional confocal
270
microscopy, can penetrate materials and provide information on their subsurface structure. To
271
illustrate this technical capability, we imaged the cuticles of M. acuminate (Banana) and X.
272
danguyi (Silver Dollar Plant; Fig. 5). M. acuminate was chosen as it is a monocot, with a visibly
273
waxy cuticle and regular files of elongated, epidermal cells. X. danguyi was chosen because of
274
its thick, glaucous cuticle composed of different waxes with different Raman spectra.
275
Three-dimensional reconstructions of confocal SRS images of M. acuminate (Fig. 5A and 5C)
276
and X. danguyi (Fig. 5B and 5C) cuticles show very different structures. The M. acuminate
277
cuticle is formed of a number of hair-like projections from the leaf surface. In the comparator
278
SEM image (Fig. 5E), these projections clump together, most likely due to the process of
279
fixation, and show a less even distribution. In contrast, the X. danguyi cuticle has an amorphous
280
surface, supported by a reticulate pattern (6 - 10 µm below the surface) of columnar structures
281
emerging from the leaf epidermis. The amorphous surface of the X. danguyi cuticle is clear on
282
the comparator SEM image (Fig. 5F), but the intriguing cuticle sub-structure is not.
283
The comparison between SRS and SEM images therefore demonstrates that sufficient contrast
284
can be produced using SRS microscopy to provide credible, structural images of waxes with
285
submicron resolution and that correlate well with images from SEM and can provide critical,
286
structural information not previously available.
287
288
Simultaneous Ultrastructural Imaging and Chemical Analysis of Cuticle by stimulated Raman
289
scattering microscopy.
290
As previously noted, SRS imaging relies on the chemically-specific Raman spectrum of the
291
cuticle constituents. To verify the potential of the technique to provide structural and chemical
292
information, simultaneously and in vivo, we first analysed the cuticle of D. anthonyi (Dudleya;
293
Fig. 6). D. anthonyi was selected for this aspect of the investigation because previous work had
294
shown stratification of the composition of cuticular wax in the D. anthonyi cuticle (Jetter et al.,
295
2000).
296
Both the SRS (Fig. 6A) and SEM (Fig. 6B) images of the D. anthonyi cuticle show a complex
297
structure; the outer cuticle layer has a structure in which many small wax crystals are amassed
298
together with the deeper layers of the cuticle having a smoother, somewhat reticulated or
299
wavelet-like appearance (Fig. 6D). Most importantly for this investigation, it is clear that the wax
300
composition, as reported by the SRS spectra (Fig. 6C) was different depending on cuticle depth.
301
The SRS spectrum acquired from deeper in the cuticle was most similar to an alkane whereas
10
302
the SRS spectrum from the surface components was markedly different, indicating a different
303
chemical composition.
304
It is well documented that changes in wax composition and deposition during development and
305
in response to abiotic stress or pathogen attack are important in the life history of many plant
306
species (Bourdenx et al 2011, Raffaele et al. 2009). T. salsuginea is particularly interesting
307
because it is a close relative of A. thaliana and T. parvula, and its cuticle composition and
308
structure alters following exposure to prolonged cold treatment or vernalisation (Teusink et al.
309
2002; Amtmann 2009; Xu et al. 2013).
310
To determine whether SRS microscopy could, in principle, document the ultrastructural and
311
chemical changes in cuticle that may occur during plant development, we imaged leaves from
312
T. salsuginea plants that had been grown either in constant temperature or exposed to a 14-day
313
long cold-shock, using SRS microcscopy and SEM (Fig. 7).
314
temperature of 20°C show a cuticle containing aggregates of waxes on the adaxial surface of
315
the leaf (Fig. 7A and 7B), and relatively little cuticular waxes on the abaxial surface (Fig. 7C and
316
7D). However, following incubation at 4˚C for 14 days, a marked alteration in cuticule structure
317
can be observed on both the adaxial and abaxial surfaces (Fig. 7C to 7H): A marked increase
318
in waxy deposits in the cuticle was observed covering the entire leaf.
319
ultrastructural aggregation of the cuticle waxes was smaller and more regular that observed on
320
the adaxial surface of leaves grown in a constant temperature of 20˚C (Fig. 7A and 7B). SRS
321
microscopy therefore precisely reported the anticipated changes in cuticle formation that occur
322
following exposure to cold in T. salsuginea, and, by extension may prove a useful, in vivo tool
323
for investigating other dynamic changes in cuticle architecture and composition during
324
development and in response to environmental or biotic perturbation.
Plants grown at a constant
Moreover, the
11
325
Discussion
326
327
The cuticle is a composite, hydrophobic layer composed mainly of waxes and cutin secreted
328
from the aerial epidermis of land plants. The cuticle performs a number of physiological roles,
329
notably acting as a barrier to water loss which, in higher plants, enables the exquisite control of
330
evapotranspiration by stomatal guard cells, increasing light reflection and heat tolerance in
331
xerophytes, and providing an essential barrier to the ingress of toxins and pathogens into the
332
plant.
333
techniques that provide only a fragmented understanding of the complexity of this essential
334
structure.
335
In this investigation, we have generated in vivo images of cuticle waxes at sub-micron
336
resolution, using spontaneous Raman scattering (SRS) microscopy. Like other forms of non-
337
linear Raman-based imaging techniques, SRS is label-free and enables the simultaneous
338
acquisition of chemically specific data and confocal quality images. Most importantly, SRS
339
microscopy relies on a change in signal intensity rather than wavelength, so images can be
340
acquired without the strong auto-fluorescent background from chloroplasts and cell walls that
341
hampers the resolution and interpretation of coherent anti-Stokes Raman scattering (CARS)
342
and spontaneous Raman images (Gierlinger et al., 2010, 2012).
343
The structures observed using SRS microscopy compared well to those seen under the SEM,
344
albeit with a lesser degree of resolution due to the inherent limitations of optical microscopy
345
relative to electron microscopy. However, unlike SEM which generates images from the surface
346
of fixed or cryogenically preserved samples, SRS imaging, like other forms of single- or multi-
347
photon laser microscopy, enables information to be acquired in vivo from within the sample and
348
three-dimensional reconstructions of internal structures to be reconstructed. This capacity is
349
particularly advantageous for documenting dynamic changes of topography or composition that
350
may occur and that are otherwise undetectable using fixed samples.
351
As demonstrated here, the chemical specificity of the SRS signal allows differentiation of
352
different wax components from the main constituents of cell walls, cellulose and pectins. In this
353
study, we showed that the xerophytes D. anthonyi and X. danguyi had different cuticular wax
354
compositions to the other, mesophytic plants investigated. Although this result is unsurprising
355
as these plants were selected for precisely their singularly thick and glaucous cuticles, it does
356
demonstrate the potential of SRS for in vivo characterisation of cuticle constituents.
357
Moreover, although we did not attempt in this study to quantify in situ the different compounds
358
for which we acquired SRS spectra, it may be possible to use this technique in a more
359
quantifiable manner to characterise, in vivo, the relative abundance of cuticle or cell wall
360
components and the changes that may occur in response to environmental or developmental
361
stimuli. In this investigation, we have shown that SRS imaging is sufficiently sensitive enough
Cuticle topography and composition has mainly been investigated using destructive
12
362
to show developmental changes in wax production in T. salsuginea. Moreover, the technique
363
may be useful for characterising wax biosynthesis mutants, such as the exemplar used here,
364
cer1, which has been shown to have a role in both water use and Arabidopsis interaction with
365
fungal and bacterial pathogens (Bourdenx et al. 2011, Raffaele et al. 2009). SRS microscopy
366
may therefore provide new insights into the physiological responses of plants to drought,
367
temperature, chemicals or pathogens, and underpin the use of alternative model systems such
368
as T. salsuginea (Amtmann 2009) to investigate the control of cuticle deposition.
369
However promising, any new technology or application must be simplified to enable widespread
370
use.
371
However, commercial, user-friendly CARS microscopes that can be converted to SRS imaging
372
exist and, like laser confocal microscopy, will become increasingly common and add real-time,
373
label-free Raman microscopy to the range of imaging techniques widely available to
374
researchers in plant science.
Currently, SRS microscopy is somewhat niche and requires specialist knowledge.
375
376
Conclusion
377
The chemical composition and structure of the cuticle in mesophyllous plants are crucial
378
elements for survival, physiology and development. We have shown that spontaneous Raman
379
scattering (SRS) microscopy can be used to acquire in-vivo, three-dimensional images of the
380
cuticle in a variety of plant species that enable simultaneous analysis of cuticle structure and
381
chemical composition. In contrast to coherent anti-Stokes Raman (CARS) microscopy, which is
382
more common, SRS images of plant tissues contain very low autofluorescence from
383
chloroplasts or cell walls and are therefore easier to interpret. Moreover, SRS images compare
384
well with those acquired using scanning electron microscopy (SEM), albeit with the lower
385
resolution that is due to the use of laser light rather than an electron beam. Moreover, we have
386
shown that SRS can be used to track and quantify dynamic changes in cuticle structure and
387
composition in response to environmental stimuli, and therefore increase our understanding of
388
this essential and often overlooked structure.
13
389
Materials and Methods
390
391
Plants.
392
Xerosicyos danguyi (Silver Dollar Plant), Dudleya anthonyi
393
(Banana), Monstera delicosa (Cheeseplant) were grown in soil, in the University of Exeter
394
glasshouses and harvested between February and April 2012. The Photoperiod was set to 16 h,
395
from 05.00 to 21.00. Supplemental lighting and shading was provided to ensure irradiance
396
beween 540 and 720 µmol s-1 m-2. The greenhouse temperature was set to 19.5 ˚C with a
397
standard deviation of 3.7 ˚C.
398
Thellungiella salsuginea and Thellungiella parvula (Saltwater Cress) were grown in a 16 h light /
399
8 h dark (16L/8D) photoperiod, at 20 ˚C, in growth rooms controlled for temperature and
400
humidity. 4 weeks after germination, Thellungiella species were transferred to an ambient
401
temperature of 4 ˚C for two weeks and then returned to 20 ˚C for a further 2 weeks. The light
402
regime remained at 16L/8D during all experiments. Leaves were harvested for imaging and
403
biochemical analysis.
(Dudleya), Musa acuminate
Arabidopsis thaliana (Mouse-eared Cress; Arabidopsis),
404
405
Purification of Wax from Plant Cuticle.
406
Following excision from plants, leaf surface areas were measured and logged. Leaf surfaces
407
were washed for 30 s in 15 ml HPLC-grade hexane (Sigma, U.K.). The hexane solvent and
408
dissolved cuticular waxes was decanted into a glass vial that had been washed with acetone
409
and dried. Leaves were washed a second time with 2 ml hexane, which was added to the
410
appropriate sample vial. Hexane was evaporated under a continuous stream of N2 to dryness
411
and the cuticle wax re-dissolved in 250 µl hexane.
412
413
Spontaneous Raman Spectroscopy.
414
Spontaneous Raman spectra of purified wax samples were acquired using a Renishaw RM100
415
Raman microscope (Renishaw plc, U.K.), equipped with a 785nm diode laser and a 1200
416
line/mm spectral grating which gave a spectral resolution of 1 cm -1. Samples were mounted on
417
aluminium-coated microscope slides.
418
419
Stimulated Raman Scattering Microscopy.
420
A detailed technical explanation of the materials and methods for stimulated Raman scattering
421
(SRS) microscopy is provided to promote the implementation of the technique as widely as
422
possible.
423
Stimulated Raman Gain Microscopy (SRG) was two laser systems were used: The Stokes
424
beam was provided by the signal output from the picosecond (ps) Optical Parametric Oscillator
425
(Levante, Emerald APE) pumped by the frequency doubled output from a Nd:Van ps oscillator
14
426
(picoTrain, HighQ laser). The pump beam was provided by a Ti:Sapphire laser (MIRA 900 D,
427
Coherent) tuned to 770 nm and operating in ps mode, which was electronically synchronised to
428
the Nd:Van laser (Coherent, Synchro-lock AP) thereby ensuring the laser pulses were
429
temporally overlapped.
430
The 770nm pump beam was amplitude modulated at 1.7MHz using an acousto-optical
431
modulator (AOM) (Crystal Technology Inc., USA, model 3080-122) and combined with the
432
signal output from the OPO using an 850 nm short pass dichroic mirror. Imaging was performed
433
using a modified confocal laser scanning unit (Flouview 300, Olympus, U.K.) and Olympus IX71
434
inverted microscope. The laser light was focused onto the sample using a 60x 1.2 / NA water
435
immersion microscope objective (UPlanS Apo, Olympus, UK). The transmitted light from the
436
sample was collected with a 60x 1.0 / NA water dipping condenser (LUMFI, Olympus, U.K.).
437
The transmitted Stokes beam was detected using a Si: photo-diode (Thorlabs, FDS1010) with a
438
70 V reverse bias. The pump beam was blocked from the photo-diode using an 850 nm long
439
pass filter (hq850lp, Chroma Technologies). The output from the photo-diode was passed
440
through a long pass filter (mini circuits, BLP-1.9+) to remove modulations at the 76 MHz laser
441
repetition rate and terminated by a 50 Ω resistor. The filtered beam was then fed into a lock-in
442
amplifier (Zurich instruments, HF2L1 Lock-in amplifier) that separated out the modulated
443
stimulated Raman scattering (SRS) signal at 1.7 MHz. SRS spectra were obtained by tuning
444
wavelength of signal beam in 0.2 nm intervals and acquiring a series of images.
445
Stimulated Raman gain was measured in preference to stimulated Raman loss, as it allowed
446
the chemically specific Raman signal to be separated from any additional signal due to two
447
photon absorption (Ye et al., 2009) or photo-thermal lensing (Moger et al., 2012; Lu et al., 2010)
448
by phase-sensitive lock-in detection (Mansfield et al., 2013).
449
Fresh, excised plant leaves were mounted in perfluorodecalin (Littlejohn et al., 2010) prior to
450
imaging, which is has been shown not to interfere with Raman-based imaging in plants
451
(Littlejohn et al. 2014; Mansfield et al. 2013).
452
453
454
Scanning Electron Microscopy
455
Leaf samples were imaged using cryogenic scanning electron microscopy. Samples were flash-
456
frozen in liquid N2 slush, transferred to a vacuum and coated in gold using the Gatan Alto 2100
457
system. Images were acquired using a JEOL JSM-6390 LV scanning electron microscope
458
operating at 5 kV with a working distance of 10-12 nm.
15
459
Acknowledgements
460
461
We thank Peter Splatt of the Exeter Bioimaging Centre for his technical assistance with the
462
scanning electron microscopy and James Chidlow for plant samples from the University of
463
Exeter glasshouse collection. Thanks also to Dr. Anna Amtmann of the University of Glasgow,
464
who generously provided seeds of Thellungiella species used. We also gratefully acknowledge
465
the support of Biomedical Physics and of the Exeter Imaging Network.
16
466
Literature Cited
467
468
AMTMANN A. 2009. Learning from Evolution: Thellungiella Generates New Knowledge on
469
Essential and Critical Components of Abiotic Stress Tolerance in Plants Mol. Plant, 2 (1): 3-12.
470
471
BAKER, E. A., & HOLLOWAY, P. J. 1971. Scanning electron microscopy of waxes on plant
472
surfaces. Micron (1971), 2(4), 364–380.
473
474
BARTHLOTT, W., NEINHUIS, C., CUTLER, D., DITSCH, F., MEUSEL, I., THEISEN, I., &
475
WILHELMI, H. 2008. Classification and terminology of plant epicuticular waxes. Botanical
476
Journal of the Linnean Society, 126(3), 237–260.
477
478
BEISSON, F., LI-BEISSON, Y., & POLLARD, M. 2012. Solving the puzzles of cutin and suberin
479
polymer biosynthesis. Current Opinion in Plant Biology, 15(3), 329–337.
480
481
BESSIRE M., CHASSOT C., JACQUAT A-C., HUMPHRY M., BOREL S., PETÉTOT J. M-C.,
482
MÉTRAUX J-P., AND NAWRATH C. 2007. A permeable cuticle in Arabidopsis leads to a strong
483
resistance to Botrytis cinerea. EMBO J., 26(8), 2158–2168.
484
485
BHUSHAN B. AND JUNG Y.C. 2006. Micro- and nanoscale characterization of hydrophobic and
486
hydrophilic leaf surfaces. Nanotechnology 17(11), 2758–2772.
487
488
BHUSHAN B., JUNG Y.C., AND KOCH K. 2009. Micro-, nano-and hierarchical structures for
489
superhydrophobicity, self-cleaning and low adhesion. Phil. Trans. R. Soc. A. 367 1631-1672.
490
491
BOURDENX B., BERNARD A., DOMERGUE F., PASCAL S., LÉGER A., ROBY D., PERVENT
492
M., VILE D., HASLAM R.P., NAPIER J.A., LESSIRE R., and JOUBÈS J. 2011 Overexpression
493
of Arabidopsis ECERIFERUM1 Promotes Wax Very-Long-Chain Alkane Biosynthesis and
494
Influences Plant Response to Biotic and Abiotic Stresses Plant Physiology 156, 29-45
495
496
BUSCHHAUS C., HAGER D., and JETTER R. 2014. Wax layers on Cosmos bipinnatus petals
497
contribute
498
10.1104/pp.114.249235
unequally
to
the
total
petal
water
resistance.
Plant
Physiology
epub
499
500
CHENG, J. X., JIA, Y. K., ZHENG, G. F. & XIE, X. S. 2002. Laser-scanning coherent anti-
501
Stokes Raman scattering microscopy and applications to cell biology. Biophysical Journal, 83,
502
502-509.
17
503
504
CHENG, J. X. & XIE, X. S. 2004. Coherent anti-Stokes Raman scattering microscopy:
505
Instrumentation, theory, and applications. Journal of Physical Chemistry B, 108, 827-840.
506
507
DELVENTHAL R., FALTER C., STRUGALA R., ZELLERHOFF N., AND SCHAFFRATH U.
508
2014 Ectoparasitic growth of Magnaporthe on barley triggers expression of the putative barley
509
wax biosynthesis gene CYP96B22 which is involved in penetration resistance., BMC Plant Biol.
510
14, 26.
511
512
DING S-H., LIU Y-S, ZENG Y, HIMMEL M.E., BAKER J.O., BAYER E.A. 2012 How Does Plant
513
Cell Wall Nanoscale Architecture Correlate with Enzymatic Digestibility? Science 338, 1055
514
515
516
EVANS, C. L. & XIE, X. S. 2008. Coherent Anti-Stokes Raman Scattering Microscopy:
517
Chemical Imaging for Biology and Medicine. Annual Review of Analytical Chemistry, 1, 883-
518
909.
519
520
FREUDIGER, C. W., MIN, W., SAAR, B. G., LU, S., HOLTOM, G. R., HE, C. W., TSAI, J. C.,
521
KANG, J. X. & XIE, X. S. 2008. Label-Free Biomedical Imaging with High Sensitivity by
522
Stimulated Raman Scattering Microscopy. Science, 322, 1857-1861.
523
524
GILBERT, R. D., JOHNSON, A. M., & DEAN, R. A. 1996. Chemical signals responsible for
525
appressorium formation in the rice blast fungusMagnaporthe grisea. Physiological and
526
Molecular Plant Pathology, 48(5), 335–346.
527
528
GREENE, P. R. & BAIN, C. D. 2005. Total internal reflection Raman spectroscopy of barley leaf
529
epicuticular waxes in vivo. Colloids and Surfaces B: Biointerfaces, 45, 174-180.
530
531
GIERLINGER N., LUSS S., KÖNIG C., KONNERTH J., EDER M., AND FRATZL P. Cellulose
532
microfibril orientation of Picea abies and its variability at the micron-level determined by Raman
533
imaging. 2010 Journal of Experimental Botany 61 (2) 587-595
534
535
GIERLINGER N., KEPLINGER T., and HARRINGTON M. Imaging of plant cell walls by
536
confocal Raman microscopy. 2012 Nature Protocols 7 (9) 1694-17-08.
537
538
GRNCAREVIC M. AND RADLER F. 1967 The effect of wax components on cuticular
539
transpiration-model experiments. Planta, 75, 23-27
18
540
541
HEREDIA-GUERRERO J.A., BENÍTEZ J.J., DOMÍNGUEZ E., BAYER I., CINGOLANI R.,
542
ATHANASSIOU A., and HEREDIA A. 2014. Infrared and raman spectroscopic features of plant
543
cuticles: a review. Frontiers in Plant Science, 5, 00305.
544
545
HEN-AVIVI S., LASHBROOKE J., COSTA F., and AHARONI A. 2014. Scratching the surface:
546
genetic regulation of cuticle assembly in fleshy fruit. J. Exp. Bot. 65(16), 4653–4664.
547
548
HO, M. & PEMBERTON, J. E. 1998. Alkyl chain conformation of octadecylsilane stationary
549
phases by Raman spectroscopy. 1. Temperature dependence. Analytical chemistry, 70, 4915-
550
4920.
551
552
JETTER, R., SCHÄFFER, S., & RIEDERER, M. 2000. Leaf cuticular waxes are arranged in
553
chemically and mechanically distinct layers: evidence from Prunus laurocerasus L. - Jetter -
554
2001 - Plant, Cell & Environment - 23, 619–628
555
556
KIM, K. W., LEE, S.-T., BAE, S.-W., & KIM, P.-G. 2011. 3D surface profiling and high resolution
557
imaging for refining the florin rings and epicuticular wax crystals of Pinus koraiensis needles.
558
Microscopy Research and Technique, 74(12), 1166–1173.
559
560
KOCH K., NEINHUIS C., ENSIKAT H-J.,
561
epicuticular waxes on living plant surfaces imaged by atomic force microscopy (AFM), J. Exp.
562
Bot. 55 (397): 711-718
and BARTHLOTT W. 2004. Self assembly of
563
564
KOCH K., BHUSHAN B., AND BARTHLOTT W. 2009. Multifunctional surface structures of
565
plants: An inspiration for biomimetics. Progress in Materials Science. 54(2), 137–178.
566
567
KOLATTUKUDY P. E. 1981 Structure, biosynthesis, and biodegradation of cutin and suberin
568
Ann. Rev. Plant Physiol. 32, 539-67
569
570
KOORNNEEF M., HANHART C.J., AND THIEL F. (1989). A Genetic and Phenotypic
571
Description of Eceriferum (cer) Mutants in Arabidopsis thaliana. Journal of Heredity 80: 118-122
572
573
KRAUSS P, MARKSTADTER C, and RIEDERER M 1997. Attenuation of UV radiation by plant
574
cuticles from woody species. Plant Cell Environ. 20(8), 1079–1085.
575
19
576
LE, T. T., YUE, S. & CHENG, J.-X. 2010. Shedding new light on lipid biology with coherent anti-
577
Stokes Raman scattering microscopy. Journal of lipid research, 51, 3091-3102.
578
579
LEE Y.H. AND DEAN R.A. (1994). Hydrophobicity of contact surface induces appressorium
580
formation in Magnaporthe grisea. FEMS Microbiology Letters. 115, 71–75
581
582
LITTLEJOHN, G. R., GOUVEIA, J. D., EDNER, C., SMIRNOFF, N., & LOVE, J. 2010.
583
Perfluorodecalin enhances in vivo confocal microscopy resolution of Arabidopsis thaliana
584
mesophyll. The New Phytologist, 186(4), 1018–1025.
585
586
LITTLEJOHN, G. R., MANSFIELD, J. C., CHRISTMAS, J. T., WITTERICK, E., FRICKER, M. D.,
587
GRANT, M. R., NICHOLAS SMIRNOFF, RICHARD M EVERSON, JULIAN MOGER, AND
588
JOHN LOVE 2014. An update: improvements in imaging perfluorocarbon-mounted plant leaves
589
with implications for studies of plant pathology, physiology, development and cell biology.
590
Frontiers in Plant Science, 5. doi:10.3389/fpls.2014.00140
591
592
LU, S., MIN, W., CHONG, S., HOLTOM, G. R. & XIE, X. S. 2010. Label-free imaging of heme
593
proteins with two-photon excited photothermal lens microscopy. Applied Physics Letters, 96,
594
113701.
595
596
LY, S., MCNERNEY, G., FORE, S., CHAN, J. & HUSER, T. 2007. Time-gated single photon
597
counting enables separation of CARS microscopy data from multiphoton-excited tissue
598
autofluorescence. Optics Express, 15, 16839-16851.
599
600
MANSFIELD, J. C., LITTLEJOHN, G. R., SEYMOUR, M. P., LIND, R. J., PERFECT, S. &
601
MOGER, J. 2013. Label-free chemically specific imaging in planta with stimulated Raman
602
scattering microscopy. Analytical Chemistry. , 85, 5055−5063
603
604
MIN W, FREUDIGER CW, LU SJ, & XIE XS (2011) Annual Review of Physical Chemistry, Vol
605
62, Annual Review of Physical Chemistry, 62, 507-530. 606
607
MOGER, J., GARRETT, N. L., BEGLEY, D., MIHOREANU, L., LALATSA, A., LOZANO, M. V.,
608
MAZZA, M., SCHATZLEIN, A. & UCHEGBU, I. 2012. Imaging cortical vasculature with
609
stimulated Raman scattering and two-photon photothermal lensing microscopy. Journal of
610
Raman Spectroscopy, 43, 668–674
611
612
20
613
NANDAKUMAR, P., KOVALEV, A. & VOLKMER, A. 2009. Vibrational imaging based on
614
stimulated Raman scattering microscopy. New Journal of Physics, 11, 033026.
615
616
OZEKI, Y., DAKE, F., KAJIYAMA, S., FUKUI, K. & ITOH, K. 2009. Analysis and experimental
617
assessment of the sensitivity of stimulated Raman scattering microscopy. Optics Express, 17,
618
3651-3658.
619
620
OZEKI Y., UMEMURA W., OTSUKA Y., SATOH S., HASHIMOTO H., SUMIMURA K.,
621
NISHIZAWA N., FUKUI K., AND ITOH K. 2012 High-speed molecular spectral imaging of tissue
622
with stimulated Raman scattering. Nature Photon 6 (12), 845-851
623
624
PIGHIN J. A., ZHENG H., BALAKSHIN L.J., GOODMAN I.P., WESTERN T. L., JETTER R.,
625
KUNST L., SAMUELS A.L. 2004. Plant Cuticular Lipid Export Requires an ABC Transporter.
626
Science, 306(5696), 702–704
627
628
RAFFAELE S, LEGER A, ROBY D 2009 Very long chain fatty acid and lipid signaling in the
629
response of plants to pathogens. Plant Signal Behav 4: 94–99
630
631
RISTIC Z. and JENKS M. A. 2002. Leaf cuticle and water loss in maize lines differing in
632
dehydration avoidance. Journal of Plant Physiology 159 (6 ) 645–651
633
SAMUELS, L., KUNST, L., & JETTER, R. 2008. Sealing plant surfaces: cuticular wax formation
634
by epidermal cells. Annual Review of Plant Biology, 59, 683–707.
635
636
SAAR, B. G., ZENG, Y. N., FREUDIGER, C. W., LIU, Y. S., HIMMEL, M. E., XIE, X. S. & DING,
637
S. Y. 2010. Label-Free, Real-Time Monitoring of Biomass Processing with Stimulated Raman
638
Scattering Microscopy. Angewandte Chemie-International Edition, 49, 5476-5479.
639
640
SNYDER, R., HSU, S. & KRIMM, S. 1978. Vibrational spectra in the C---H stretching region and
641
the structure of the polymethylene chain. Spectrochimica Acta Part A: Molecular Spectroscopy,
642
34, 395-406.
643
644
VERAVERBEKE, E. A., VAN BRUAENE, N., & VAN OOSTVELDT, P. 2001. Non destructive
645
analysis of the wax layer of apple (Malus domestica Borkh.) by means of confocal laser
646
scanning microscopy Planta, 213, 525-533.
647
21
648
WEISSFLOG, I., VOGLER, N., AKIMOV, D., DELLITH, A., SCHACHTSCHABEL, D., SVATOS,
649
A., BOLAND, W., DIETZEK, B. & POPP, J. 2010. Toward in Vivo Chemical Imaging of
650
Epicuticular Waxes. Plant Physiology, 154, 604.
651
652
XU X, FENG J, LÜ S, LOHREY GT, AN H, ZHOU Y, JENKS MA. 2013 Leaf cuticular lipids on
653
the Shandong and Yukon ecotypes of saltwater cress, Eutrema salsugineum, and their
654
response to water deficiency and impact on cuticle permeability. Physiologia Plantarum 446–
655
458.
656
657
YE, T., FU, D. & WARREN, W. S. 2009. Nonlinear Absorption Microscopy†. Photochemistry
658
and Photobiology, 85, 631-645.
659
660
YEATS, T. H., & ROSE, J. K. C. 2013. The Formation and Function of Plant Cuticles. Plant
661
Physiology. 163, 5–20
662
663
YU, M. M. L., KONOROV, S. O., SCHULZE, H. G., BLADES, M. W., TURNER, R. F. B., and
664
JETTER, R. (2008). In situ analysis by microspectroscopy reveals triterpenoid compositional
665
patterns within leaf cuticles of Prunus laurocerasus. Planta, 227(4), 823–834.
666
667
YU Y., RAMACHANDRAN P.V., and WANG M.C. 2014 Shedding new light on lipid functions
668
with CARS and SRS microscopy., Biochem Biophys Acta 1841(8), 1120–1129.
669
670
ZENG, Y., SAAR, B., FRIEDRICH, M., CHEN, F., LIU, Y.-S., DIXON, R., HIMMEL, M., XIE, X. &
671
DING, S.-Y. 2010. Imaging Lignin-Downregulated Alfalfa Using Coherent Anti-Stokes Raman
672
Scattering Microscopy. BioEnergy Research, 3, 272-277.
673
674
ZUMBUSCH, A., HOLTOM, G. R. & XIE, X. S. 1999. Three-dimensional vibrational imaging by
675
coherent anti-Stokes Raman scattering. Physical Review Letters, 82, 4142-4145.
676
677
678
679
680
Figure and Table Legends
681
682
Figure 1 Title: Schematic representation of the two coherent Raman scattering processes:
683
coherent anti-Stokes Raman scattering (CARS) and stimulated Raman scattering (SRS).
684
22
685
Figure 1 Legend: Panel A: Energy level diagrams for the CARS and SRS processes, showing
686
the pump (green), Stokes (red) and anti-Stokes (blue) photon energies,
687
Panel B: Diagrammatic representation of the input and output spectra for CARS and SRS,
688
showing the gain and loss in the pump (red) and Stokes (green) beams respectively.
689
Panel C: Diagrammatic representation of the modulation transfer detection scheme used to
690
detect stimulated Raman gain and loss with high sensitivity.
691
692
693
694
Figure 2 Title: Raman spectra and curve fits of plant leaf waxes.
695
696
Figure 2 Legend: A: Raman spectra of cuticle waxes purified from Xerosicyos danguyi (Silver
697
Dollar Plant), Dudleya anthonyi
698
Arabidopsis), Musa acuminate (Banana) and Thellungiella parvula (Saltwater Cress). Peak
699
fitting to the Raman spectra of T. parvula (B), D. antonyi (C) and X. danguyi (D).
(Dudleya) Arabidopsis thaliana (Mouse-Eared Cress;
700
701
702
703
Figure 3 Title:
704
images of the Thellungelia parvula leaf surface.
Coherent anti-Stokes Raman scattering and stimulated Raman scattering
705
706
Figure 3 Legend: The surface of a leaf from a Thellungelia parvula was simultaneously imaged
707
using coherent anti-Stokes Raman scattering (CARS) microscopy (panel A) and stimulated
708
Raman scattering (SRS) microscopy (panel B). Both images were acquired at 2845cm -1 CH2
709
symmetric stretch. The CARS image (A) is dominated by autofluorescence, and only the largest
710
wax crystals are visible against the background. Conversely, the SRS image (B) is almost
711
background free, enabling the cell walls and wax crystals to be clearly visualised. Panel C
712
shows the spectral scan of the SRS (red crosses) from the wax crystals in situ overlaid onto the
713
spontaneous Raman spectra of purified T. parvula wax (blue line), purified cellulose (green line)
714
and purified pectin (purple line).
715
716
717
718
Figure 4 Title: Stimulated Raman scattering images and scanning electron micrographs of
719
Arabidopsis thaliana cuticle.
720
23
721
Figure 4 Legend: The surface of Arabidopsis thaliana stems were imaged using stimulated
722
Raman scattering microscopy (panels A, C and D) and scanning electron microscopy (panels,
723
B, D and E).
724
ecotype “Landsberg erecta”, with the structure of cuticle wax crystals clearly visible. Panels C
725
and D show the surface of the wild-type A. thaliana stem following a wash in hexane that
726
dissolved the cuticle waxes. Panels E and F show the surface of the stem from a cer 1 A.
727
thaliana mutant. Images A, C and E are generated from 3D stacks taken of a 64 x 64 µm field of
728
view and are displayed in false colour.
Panels A and B show the surface of the untreated stem of wild-type A thaliana,
729
730
731
732
Figure 5 Title: Stimulated Raman scattering images and scanning electron micrographs of
733
cuticle of Musa acuminate and of Xerosicyos danguyi.
734
735
Figure 5 Legend: The surface of Musa acuminate (banana) and of Xerosicyos danguyi (Silver
736
dollar Plant) leaves were imaged using stimulated Raman scattering (SRS) microscopy (panels
737
A, B, C and D) and scanning electron microscopy (SEM) (panels E and F). Panel A is a SRS
738
image of the adaxial leaf surface of surface M. acuminate constructed from an image stack from
739
a 64 x 64 µm field of view. Panel B is a SRS image of X. danguyi leaf reconstructed from an
740
image stack from a 126 x 126 µm field of view. Panels C and D are orthogonal views projected
741
from the image stacks presented in A and B, respectively, and show the depth profile of
742
cuticles. Panels E and F are SEM images of the cuticles of M. acuminate and X. danguyi
743
leaves, respectively.
744
745
746
747
Figure 6 Title: Stimulated Raman scattering spectral images of Dudleya anthonyi cuticle.
748
749
Figure 6 Legend: Stimulated Raman scattering (SRS) spectral imaging of Dudleya anthonyi
750
(Dudleya) cuticle waxes shows changes in chemical composition with depth. Panels A and B
751
are, respectively, the SRS and SEM images of D. anthonyi leaf showing the crystalline structure
752
of cuticle wax deposits on the adaxial surface. The image in panel A is a 3-D reconstruction of
753
an image stack taken from a 126 x 126 µm field of view at the 2840 cm -1 Raman shift. Panel C
754
shows the SRS spectral scans taken from the more superficial and deeper wax areas, indicated
755
on Panel D. The blue rectangle indicates an area of superficial wax and the red line delimits an
756
area deeper in the cuticle. The panels on the right and lower sides of the main image in Panel
24
757
D are from the orthogonal views through the stack showing the depth profile image of the cuticle
758
wax.
759
760
761
762
Figure 7 Title: Stimulated Raman scattering images and scanning electron micrographs of
763
cuticle of the cuticle of Thellungiella salsuginea following cold-induced wax biogenesis. Musa
764
765
Figure 7 Legend: Thellungiella salsuginea (Saltwater Cress) leaves were imaged using
766
Stimulated Raman scattering (SRS) spectral imaging and scanning electron microscopy (SEM)
767
before and after cold treatment. Panels A, C, E and G are three dimensional reconstructions of
768
image stacks from a 250 x 250 µm field of view, at the 2840cm -1 Raman shift. Panels B, D, F,
769
and H are SEM images; scalebars represent 50 µm. Panels A-B and C-D are, respectively, from
770
the adaxial and abaxial leaf surfaces of plants grown at 20°C for 8 weeks. Panels E-F and G-H
771
are, respectively, from adaxial and abaxial leaf surfaces of plants grown at 20°C for 4 weeks,
772
then at 4˚C for 2 weeks and at 20 ˚C for a further 2 weeks.
773
774
775
776
Table 1 Title: Raman peaks used in curve fitting for cuticular waxes.
25
Download