1. Trang chủ
  2. » Giáo án - Bài giảng

breath figures on leaf surfaces formation and effects of microscopic leaf wetness

31 1 0

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

THÔNG TIN TÀI LIỆU

IN PLANT NUTRITION ‘Breath figures’ on leaf surfaces – formation and effects of microscopic leaf wetness Jürgen Burkhardt and Mauricio Hunsche Journal Name: Frontiers in Plant Science ISSN: 1664-462X Article type: Hypothesis & Theory Article Received on: 01 Jul 2013 Accepted on: 04 Oct 2013 Provisional PDF published on: 04 Oct 2013 Frontiers website link: www.frontiersin.org Citation: Burkhardt J and Hunsche M(2013) ‘Breath figures’ on leaf surfaces – formation and effects of microscopic leaf wetness 4:422 doi:10.3389/fpls.2013.00422 Article URL: http://www.frontiersin.org/Journal/Abstract.aspx?s=1220& name=plant%20nutrition&ART_DOI=10.3389/fpls.2013.00422 (If clicking on the link doesn't work, try copying and pasting it into your browser.) Copyright statement: © 2013 Burkhardt and Hunsche This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY) The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice No use, distribution or reproduction is permitted which does not comply with these terms This Provisional PDF corresponds to the article as it appeared upon acceptance, after rigorous peer-review Fully formatted PDF and full text (HTML) versions will be made available soon 1    1  ‘Breath figures’ on leaf surfaces – formation and effects of 2  microscopic leaf wetness 3  Juergen Burkhardt1 & M Hunsche2 4  5  6  Revision 7  8  9  10  11  University of Bonn, Karlrobert-Kreiten-Str 13 12  D-53115 Bonn, Germany Institute of Crop Science and Resource Conservation, Plant Nutrition Group 13  14  15  University of Bonn, Auf dem Hügel 16  D-53121 Bonn, Germany Institute of Crop Science and Resource Conservation, Horticultural Science Group 17  18  Corresponding author: 19  Juergen Burkhardt, Tel +49 228 732186, email: j.burkhardt@uni-bonn.de, 20  21  22  Submitted to Frontiers in Plant Nutrition       2    23  Abstract 24  ‘Microscopic leaf wetness’ means minute amounts of persistent liquid water on leaf surfaces which are 25  invisible to the naked eye The water is mainly maintained by transpired water vapor condensing onto 26  the leaf surface and to attached leaf surface particles With an estimated average thickness of less than 27  µm, microscopic leaf wetness it is about orders of magnitude thinner than morning dewfall 28  The most important physical processes which reduce the saturation vapor pressure and promote 29  condensation are cuticular absorption and the deliquescence of hygroscopic leaf surface particles 30  Deliquescent salts form highly concentrated solutions Depending on the amount and concentration of 31  the dissolved ions, the physicochemical properties of microscopic leaf wetness can be considerably 32  different from those of pure water Microscopic leaf wetness can form continuous thin layers on 33  hydrophobic leaf surfaces and in specific cases can act similar to surfactants, enabling a strong 34  potential influence on the foliar exchange of ions Microscopic leaf wetness can also enhance the 35  dissolution, the emission, and the reaction of specific atmospheric trace gases e.g ammonia, SO2, or 36  ozone, leading to a strong potential role for microscopic leaf wetness in plant/atmosphere interaction 37  Due to its difficult detection, there is little knowledge about the occurrence and the properties of 38  microscopic leaf wetness However, based on the existing evidence and on physicochemical reasoning 39  it can be hypothesized that microscopic leaf wetness occurs on almost any plant worldwide and often 40  permanently, and that it significantly influences the exchange processes of the leaf surface with its 41  neighboring compartments, i.e., the plant interior and the atmosphere The omission of microscopic 42  water in general leaf wetness concepts has caused far-reaching, misleading conclusions in the past 43  44  Keywords: 45  aerosols, cloud condensation nuclei, contact angle, deliquescence point, dew, foliar nutrition, 46  Hofmeister series, leaf surfaces, leaf boundary layer, plant-atmosphere interaction, surface tension, 47  stomatal uptake       3    48  Abbreviations AMS: ammonium sulphate, CET: Central European Time, DRH: deliquescence 49  relative humidity, ESEM: environmental scanning electron microscope, HAS: hydraulic activation of 50  stomata, LBL: Leaf boundary layer, PAR: photosynthetically active radiation, RH: relative humidity, 51  RHs: relative humidity at the leaf surface, SEM: scanning electron microscopy 52  53  54  55  56  57  58  59  60  61  62  63  64  65        4    66  Introduction 67  ‘Breath figures‘ is a term used in material science to describe the condensation as well as the linked 68  wetting and dewetting processes on different kinds of surfaces (Blaschke et al., 2012) The 69  examination of breath figures has then been used as a method to characterize the degree of 70  contamination on an otherwise homogenous surface (Kumar and Whitesides, 1994) The term was 71  originally introduced by Aitken (1914) who noticed that water from exhaled breath condensing to 72  clean glass surfaces was clearly visible as separate droplets If the glass was contaminated with fine 73  particles, however, the condensation would be strong but not visible, due to the formation of thin 74  water films (Aitken, 1911) Condensation to deposited particles (‘contaminants’) is also considered an 75  essential factor in corrosion, and according to ISO 9223 wetting happens at 80% RH and above due to 76  particle hygroscopicity (Schindelholz and Kelly, 2012) 77  In plant science, the influence of particles on condensation has not been considered sufficiently so far 78  On leaf surfaces, the commonly known form of condensation is morning dewfall It develops during 79  clear, calm nights, when plant surfaces cool down by radiational heat loss, and the surface temperature 80  eventually reaches the dew point of the surrounding air According to this common meteorological 81  definition, dew formation thus starts when 100% relative humidity (RH) is reached at the actual leaf 82  surface temperature, which normally means about 90% RH of the surrounding air (Monteith, 1957) It 83  is usually neglected that the initiation of condensation on leaf surfaces likely starts on condensation 84  nuclei, analogously to atmospheric cloud formation (Beysens, 1995) These nuclei are tiny 85  hygroscopic particles, which are present on all kinds of leaf surfaces They result from atmospheric 86  dry deposition of aerosols or residues from evaporated rain droplets, while removal by rain is never 87  complete (Neinhuis and Barthlott, 1998; Freer-Smith et al., 2005) Almost all aerosols are (partly) 88  hygroscopic (Pöschl, 2005) and therefore cause a local reduction of the saturation vapor pressure 89  Even the commonly used expression ‘dry deposition’ for aerosols is usually misleading, because many 90  of the deposited substances become deliquescent at higher humidities (e.g., 75 % RH for a NaCl       5    91  particle) Equilibration with the surrounding RH happens very quickly (Pilinis et al., 1989) and many 92  particles will therefore reach a transpiring leaf surface in deliquescent form 93  Neglecting particle deliquescence can cause misleading conclusions An example is the ‘wax 94  degradation’ phenomenon that was frequently found on conifer needles which were affected by air 95  pollution caused forest decline The phenomenon was intensively investigated in the 1980s and 90s, 96  but the investigations concentrated on the chemical composition of the waxes and could not explain 97  the development of the phenomenon However, the characteristic, amorphous appearance of 98  epicuticular waxes can also be produced in a simple way by deliquescent particles covering the 99  structures of the epicuticular waxes This alternative explanation was suggested recently (Burkhardt, 100  2010) and its capability to explain the phenomenon was meanwhile demonstrated by experiment 101  (Burkhardt and Pariyar, 2013) Because the minimum epidermal conductance gmin, a key factor of tree 102  drought tolerance, was also reduced by salt particles, and given the fact that particle accumulation on 103  conifers can reach the amount of leaf waxes (up to more than 50 µg cm-2, Saebo et al., 2012), a direct 104  link between particulate air pollution and drought symptoms of conifers might exist, with ‘wax 105  degradation’ as an indication of particle load (Burkhardt and Pariyar, 2013) 106  The second neglected factor for the formation of leaf wetness is foliar (mainly stomatal) transpiration 107  In the common definition of dewfall, the main source of water vapor for dew formation on plants is 108  the surrounding atmosphere, with an eventual contribution by ‘distillation’ from the soil (Monteith, 109  1957) On leaf surfaces, however, foliar transpiration is an additional water vapor source The leaf 110  boundary layer is humidified by this water vapor, leading to high water vapor concentration especially 111  at the leaf surface (Schuepp, 1993; Roth-Nebelsick, 2007), which together with hygroscopic 112  substances will lead to the formation of microscopic leaf wetness (Burkhardt and Eiden, 1994; 113  Burkhardt et al., 1999) Although this process only involves small amounts of water, it might 114  considerably change the transport between the leaf surface and the neighboring compartments, which 115  is supported by the dependence of trace gas deposition on RH: For easily soluble compounds like NH3 116  and SO2, increasing trace gas deposition to cuticular surfaces (‘non-stomatal fluxes’) was already 117  found for 70% RH (van Hove et al., 1989;Burkhardt and Eiden, 1994;Wichink Kruit et al., 2008) The       6    118  trace gas deposition to microscopic leaf wetness is also dependent on the chemical composition of the 119  water, e.g on pH or on leached manganese ions catalyzing SO2 oxidation (Burkhardt and Drechsel, 120  1997) Non-stomatal deposition is also significant for ozone, making up between 1/3 and 2/3 of total 121  deposition (Coyle et al., 2009;Fowler et al., 2009; Launiainen et al., 2013) A positive relation of 122  ozone deposition with relative humidity was also found (Pleijel et al., 1995, Altimir et al., 2006 123  Lamaud et al., 2009) 124  Foliar fertilization is a complicated process with foliar uptake being the first decisive step (Fernandez 125  and Brow, 2013) Continuing microscopic leaf wetness might contribute considerably to the foliar 126  exchange of ions When dilute solutions are applied, the highest uptake rates into leaves occur during 127  the drying phase, presumably as a consequence of increasing concentrations (Eichert and Burkhardt, 128  2001) The high concentrations of electrolytes in deliquescent particles are expected to promote the 129  gradient dependent exchange process across the leaf surface, and maintenance of high concentrations 130  would therefore lead to high transport rates 131  Macroscopic leaf wetness, i.e visible wetting of leaves, usually has a large influence on the 132  phyllosphere For phyllospheric organisms, water is a key issue to survive (Beattie, 2011; Vorholt, 133  2012) The amount of water needed depends on the organism but usually ‘free water’ (probably 134  meaning visible water) is required by phyllospheric organisms like fungi, bacteria or insects and thus 135  fosters phyllospheric life including plant pathogens (Huber and Gillespie, 1992) Microscopic leaf 136  wetness might also influence the phyllosphere to a certain degree, but cannot be treated here in depth 137  The aim of this contribution is to elucidate the mechanisms and conditions by which microscopic leaf 138  wetness is formed and maintained So far there have only been isolated reports and phenomenological 139  descriptions, while an integrated view and a general concept detailing the occurrence and the functions 140  of microscopic liquid water at the plant/atmosphere interface is missing 141  142        7    143  Detection of microscopic leaf wetness 144  The most common method to determine (macroscopic) leaf wetness duration is the electrical resistance 145  measurement of artificial leaves A continuous resistance signal is produced, which is divided into 146  ‘wet’ or ‘dry’ by defining a resistance threshold, based on the visual observation of wetness (Gillespie 147  and Kidd, 1978;Fuentes and Gillespie, 1992;Huber and Gillespie, 1992;Armstrong et al., 148  1993;Sentelhas et al., 2007) For the detection of microscopic leaf wetness, a similar electronic device 149  can be used, but the sensors to measure the electric resistance are directly attached to the leaf surface 150  (Burkhardt and Gerchau, 1994) The signal is then compared to ambient RH (Burkhardt and Eiden, 151  1994), or to the signal of a commercial leaf wetness sensor, i.e an artificial leaf An example for the 152  latter procedure is shown in Figure The electrical conductance on potato leaves was measured in 153  Southern Germany during a hot summer week, and was compared to the continuous signal of an 154  artificial leaf sensor (237 Leaf Wetness Sensing Grid, Campbell Scientific, Logan, UT, USA) which 155  was installed in close proximity Photosynthetically active radiation (PAR) and ambient relative 156  humidity (RH) data were obtained from a weather station on the same field For both wetness sensors, 157  the nighttime increase is clearly visible and goes parallel with each other, with a significant decrease 158  of resistance starting at about 60 to 70% RH of the surrounding air During daytime, a different course 159  of the signals is observed, with the sensor on the potato leaves showing a regular increase in the 160  mornings, which is missing on the artificial leaf 161  162  Insert Fig here 163  164  Because the leaf wetness signal is highly correlated with PAR, it is most probably the consequence of 165  changing stomatal conductance, where transpired water coming from the stomata re-condenses on the 166  leaf surface This interpretation is supported by the results of a detailed study under completely 167  controlled conditions using the same type of leaf wetness sensors on bean leaves Under constant       8    168  humidity and by changing light or changing CO2 concentration was the electrical leaf surface 169  conductance closely correlated with stomatal conductance (Burkhardt et al., 1999) These results 170  indicate that microscopic water can exist on leaf surfaces for extended times, even under hot, dry 171  summertime conditions, and that the liquid water therefore is in an equilibrium state, reacting quickly 172  to increased transpiration by the formation of more liquid water, and by a reduction of the water 173  amount when the stomata close This phenomenon can be explained by two processes One is the leaf 174  transpiration which creates a humid leaf boundary layer (LBL) including the proper leaf surface 175  During times of open stomata, leaf surface humidity (RHs) will mostly be determined by transpiration, 176  with only limited influence by ambient RH The distribution of leaf surface humidity is heterogeneous 177  and will especially be high near to stomata (Schuepp, 1993;Roth-Nebelsick, 2007).The second process 178  is a local reduction of the saturation vapor pressure by effects of the leaf surface material (sorption by 179  the cuticle, deliquescence of hygrocopic leaf surface particles), or geometry (capillary condensation), 180  which will be discussed in more detail within the next section As also calculated in the next section, 181  the hypothetical homogenous thickness of the liquid water is less than µm This small amount of 182  water is not visible and is two orders of magnitude smaller than normal morning dewfall of up to 0.5 183  mm (Monteith, 1957) Although microscopic leaf wetness could be interpreted as a specific form of 184  dewfall, meteorological instruments are not sensitive enough to detect and to filter it from other 185  signals, neither by lysimeters for the amount of water, nor by flux measurements for the contribution 186  to the energy budget 187  So far, no field measurement techniques are known other than the indirect method where the signals 188  from leaf wetness sensors are compared to ambient relative humidity or to the signals from artificial 189  leaf wetness sensors Microscopic leaf wetness is also not visible without the use of microscopic 190  techniques While a combination of a gas exchange cuvette with a light microscope enabled the 191  observation of microscopic water formed by stomatal transpiration and showed the influence of the 192  leaf boundary layer (Burkhardt et al., 2001), the resolution of a light microscope is not high enough to 193  study the interactions between leaf surface particles and stomatal transpiration Detailed observations 194  are enabled by environmental scanning electron microscopy (ESEM), where it is possible to study       9    195  condensation processes at high resolution and under controlled humidity A limitation of the ESEM 196  technique to keep in mind is the fact that leaves are abscised and are not transpiring anymore, so RH 197  and RHs are only regulated from outside Another difficulty is the exact detection of leaf surface 198  temperature in case thicker leaves or needles are used, because the necessary cooling happens from a 199  small table below the sample The ESEM observations are usually done at low temperatures of 2°C to 200  5°C in order to reduce the necessary amount of water vapor molecules to reach high RH, which in 201  most cases is not a limitation ESEM observations have been used to study both condensation on 202  ambient, untreated leaves and the changes resulting from changes in RH after spraying leaves with 203  different types of solutions or dry aerosols (Burkhardt et al., 2012; Burkhardt and Pariyar, 2013) 204  205  206  207  208  209  210  211  212  213  214        16    370  371  Insert Movie here 372  373  374  It is important to note that both movies not show transpiration effects, as needles were abscised and 375  were within the vacuum chamber of the ESEM RH was only manipulated from outside It also has to 376  be noted that the ‘stomatal openings’ only show the entrance to the epistomatal chamber of the pine 377  needles The guard cells are located at the bottom of this opening and cannot be seen Nevertheless, 378  regarding the geometrical situation of interest, the epistomatal chamber has the same features as an 379  open stoma, i.e., a diverging and a converging portion This makes it comparable to the geometrical 380  situation used by (Schönherr and Bukovac, 1972) to derive their conclusion that water uptake into the 381  stomata is impossible 382  In both movies, the strong dynamics of deliquescence can be seen Movie shows the repeated 383  deliquescence and efflorescence of KI The efflorescence of the KI crystals is highly unpredictable and 384  repeatedly the crystallization takes place within the epistomatal chambers, a clear indication that KI 385  solution had entered there The movement of the solution into epistomatal chambers can be seen even 386  clearer in Movie Here, KSCN was used because it is on the far chaotropic side of the Hofmeister 387  series The movies follow one deliquescence process of KSCN The solution shows an extremely flat 388  contact angle, and it is clearly recognizable that the deliquescent KSCN solution enters the epistomatal 389  chamber The movies can thus be taken as an additional proof for stomatal uptake of aqueous 390  solutions They can also be interpreted as a first successful support for the hypothesis that chaotropic 391  salts are more easily penetrating into the stomata Finally, they can be taken as a confirmation of 392  Aitken’s observation of ‘breath figures’, i.e liquid water in a flat, non-droplet like shape on a 393  hydrophobic surface, which is spreading out easily and forming a thin water layer on the leaf surface 394        17    395  Conclusions and recommendations 396  Microscopic leaf wetness can play an important role for trace gas deposition and for ion fluxes across 397  the plant surface Increased ammonia deposition over a Douglas fir forest was observed above 70 % 398  RH at night and even lower at daytime (Wyers and Erisman, 1998), and over a grassland above 71 % 399  RH (Wichink Kruit et al., 2008) During daytime, a contribution of 66% to 88% was found for 400  ‘cuticular ammonia deposition’ to a maize canopy (Walker et al., 2013) For ammonia, this 401  microscopic leaf wetness will enable bi-directional ‘cuticular’ gaseous exchange, depending on 402  dynamic environmental conditions and the respective compensation points for different trace gases 403  (Flechard et al., 1999;Burkhardt et al., 2009;Sutton et al., 2009) Non-stomatal ozone deposition is 404  more difficult to explain, as ozone is less soluble than ammonia, and no obvious chemical reactions 405  can account for the observed non-stomatal losses However, several reaction mechanisms of ozone 406  with atmospheric aerosols have been discussed (Jacob, 2000;Oum et al., 2003;Roeselova, 2003), and 407  although such mechanisms have so far been out of focus in the search for reasons explaining non- 408  stomatal ozone deposition, they should be re-considered taking into account the likely continuing 409  occurrence of highly concentrated solutions on leaf surfaces 410  Microscopic leaf wetness influences plant physiology Leaf surface particles increase HAS, and the 411  liquid water connections formed between the leaf surface and the apoplast along the stomatal walls 412  have an influence on water and nutrient fluxes Increased transpiration and reduced water use 413  efficiency caused by leaf surface particles was observed for particle exclusion (Pariyar et al., 2013) as 414  well as for particle amendment (Burkhardt et al., 2001) The stomatal uptake of nutrients is enabled as 415  well as the stomatal leaching of ions, although an experimental proof for the latter is still missing 416  Sound reasons for nocturnal transpiration (Caird et al., 2007) have so far been missing, and nocturnal 417  stomatal nutrient uptake might represent one benefit for the plant 418  The development of models addressing both the physical mechanisms as well as the 419  (physico)chemistry of microscopic leaf wetness would be useful So far, morning dewfall is 420  considered a micrometeorological phenomenon and is assessed via a negative energy balance In order       18    421  to address the relevance of the mechanism, the implementation of microphysical aerosol models would 422  be useful, introducing ‘DCN’ (dew condensation nuclei) on leaf surfaces, with a similar formalism as 423  atmospheric CCN For this purpose, advanced chemical aerosol models could be introduced into 424  models of plant-atmosphere interaction 425  The influence of deposited aerosols on plant physiology and on plant-atmosphere interactions has so 426  far been neglected in plant science as well as in micrometeorology Leaf surface particles were 427  assumed to stay chemically inert Leaf surface wetness was defined by visible detection and was 428  considered to exist as pure water or strongly dilute solutions Microscopic leaf wetness develops by 429  the hygroscopic action of fine particles, with water vapor mainly from stomatal transpiration ‘Breath 430  figures’ on leaf surfaces are microscopically thin films as well as droplets, which are highly dynamic 431  in concentration and extension They interact with the atmosphere by bi-directional gas fluxes and 432  with the apoplast via HAS by hydraulic signals and the exchange of aqueous solutions The 433  consideration of these processes in broadened concepts of plant-atmosphere interactions is highly 434  desirable Including existing aerosol models into leaf surface exchange models seems a priority task 435  on this road 436  437  Acknowledgments 438  The authors thank Knut Wichterich for his support in the ESEM This work was supported by a a 439  research grant (BU 1099/7-1) from the Deutsche Forschungsgemeinschaft (DFG), which is gratefully 440  acknowledged 441  442  443        19    444  References: 445  Aitken, J. (1911). Breath figures. Nature 86, 516‐517.  446  Altimir, N., Kolari, P., Tuovinen, J.P., Vesala, T., Back, J., Suni, T., Kulmala, M., and Hari, P. (2006).  447  Foliage surface ozone deposition: a role for surface moisture? Biogeosciences 3, 209‐228.  448  Armstrong, R., Barthakur, N.N., and Norris, E. (1993). A comparative study of 3 leaf wetness  449  450  sensors. International Journal of Biometeorology 37, 7‐10.  Aryal, B., and Neuner, G. (2010). Leaf wettability decreases along an extreme altitudinal gradient.  451  452  Oecologia 162, 1‐9.  Beattie, G.A. (2011). "Water Relations in the Interaction of Foliar Bacterial Pathogens with Plants,"  453  in Annual Review of Phytopathology, Vol 49, eds. N.K. Vanalfen, G. Bruening & J.E. Leach.   454  (Palo Alto: Annual Reviews), 533‐555.  455  Beysens, D. (1995). The Formation of Dew. Atmospheric Research 39, 215‐237.  456  Blaschke, J., Lapp, T., Hof, B., and Vollmer, J. (2012). Breath Figures: Nucleation, Growth,  457  458  Coalescence, and the Size Distribution of Droplets. Physical Review Letters 109.  Bostrom, M., Williams, D.R.M., and Ninham, B.W. (2001). Surface tension of electrolytes: Specific  459  460  ion effects explained by dispersion forces. Langmuir 17, 4475‐4478.  Boyce, R.L., Mccune, D.C., and Berlyn, G.P. (1991). A Comparison of Foliar Wettability of Red  461  Spruce and Balsam Fir Growing at High Elevation. New Phytologist 117, 543‐555.  462  Burkhardt, J. (2010). Hygroscopic particles on leaves: Nutrients or desiccants? Ecological  463  Monographs 80, 369‐399.  464  Burkhardt, J., Basi, S., Pariyar, S., and Hunsche, M. (2012). Stomatal penetration by aqueous  465  solutions ‐ an update involving leaf surface particles. New Phytologist 196, 774‐787.  466  Burkhardt, J., and Drechsel, P. (1997). The synergism between SO2 oxidation and manganese  467  leaching on spruce needles ‐ A chamber experiment. Environmental Pollution 95, 1‐11.        20    468  Burkhardt, J., and Eiden, R. (1994). Thin water films on coniferous needles. Atmospheric  469  470  Environment 28, 2001‐2011.  Burkhardt, J., Flechard, C.R., Gresens, F., Mattsson, M., Jongejan, P.a.C., Erisman, J.W., Weidinger,  471  T., Meszaros, R., Nemitz, E., and Sutton, M.A. (2009). Modelling the dynamic chemical  472  interactions of atmospheric ammonia with leaf surface wetness in a managed grassland  473  canopy. Biogeosciences 6, 67‐83.  474  Burkhardt, J., and Gerchau, J. (1994). A new device for the study of water‐vapor condensation and  475  gaseous deposition to plant‐surfaces and particle samples. Atmospheric Environment 28,  476  2012‐2017.  477  Burkhardt, J., Kaiser, H., Goldbach, H., and Kappen, L. (1999). Measurements of electrical leaf  478  surface conductance reveal recondensation of transpired water vapour on leaf surfaces.  479  Plant Cell and Environment 22, 189‐196.  480  Burkhardt, J., Kaiser, H., Kappen, L., and Goldbach, H.E. (2001). The possible role of aerosols on  481  stomatal conductivity for water vapour. Basic and Applied Ecology 2, 351‐364.  482  Burkhardt, J., and Pariyar, S. (2013). Particulate pollutants are capable to ‘degrade’ epicuticular  483  waxes and to decrease the drought tolerance of Scots pine (Pinus sylvestris L. ).  484  Environmental Pollution, doi: 10.1016/j.envpol.2013.04.041  485  Caird, M.A., Richards, J.H., and Donovan, L.A. (2007). Nighttime stomatal conductance and  486  487  transpiration in C‐3 and C‐4 plants. Plant Physiology 143, 4‐10.  Cape, J.N. (1983). Contact angles of water droplets on needles of Scots pine (Pinus sylvestris)  growing in polluted atmospheres. New Phytologist 93, 293‐299.  488  489  Chamel, A., Pineri, M., and Escoubes, M. (1991). Quantitative determination of water sorption by  490  491  plant cuticles. Plant Cell and Environment 14, 87‐95.  Collins, K.D., and Washabaugh, M.W. (1985). The Hofmeister effect and the behavior of water at  492  interfaces. Quarterly Reviews of Biophysics 18, 323‐422.        21    493  Coyle, M., Nemitz, E., Storeton‐West, R., Fowler, D., and Cape, J.N. (2009). Measurements of ozone  494  deposition to a potato canopy. Agricultural and Forest Meteorology 149, 655‐666.  495  Deegan, R.D., Bakajin, O., Dupont, T.F., Huber, G., Nagel, S.R., and Witten, T.A. (1997). Capillary  496  flow as the cause of ring stains from dried liquid drops. Nature 389, 827‐829.  497  Dominguez, E., Heredia‐Guerrero, J.A., and Heredia, A. (2011). The biophysical design of plant  498  499  cuticles: an overview. New Phytologist 189, 938‐949.  dos Santos, A.P., Diehl, A., and Levin, Y. (2010). Surface Tensions, Surface Potentials, and the  500  501  Hofmeister Series of Electrolyte Solutions. Langmuir 26, 10778‐10783.  Dutcher, C.S., Wexler, A.S., and Clegg, S.L. (2010). Surface Tensions of Inorganic Multicomponent  502  Aqueous Electrolyte Solutions and Melts. Journal of Physical Chemistry A 114, 12216‐ 503  12230.  504  Eichert, T., and Burkhardt, J. (2001). Quantification of stomatal uptake of ionic solutes using a new  505  506  model system. Journal of Experimental Botany 52, 771‐781.  Eichert, T., Kurtz, A., Steiner, U., and Goldbach, H.E. (2008). Size exclusion limits and lateral  507  heterogeneity of the stomatal foliar uptake pathway for aqueous solutes and water‐ 508  suspended nanoparticles. Physiologia Plantarum 134, 151‐160.  509  Eiden, R., Burkhardt, J., and Burkhardt, O. (1994). Atmospheric aerosol particles and their role in  510  511  the formation of dew on the surface of plant leaves. Journal of Aerosol Science 25, 367‐376.  Fernandez, V., and Brown, P. (2013). From plant surface to plant metabolism: the uncertain fate of  512  513  foliar‐applied nutrients. Frontiers in Plant Science 4, 00289.  Fernandez, V., and Eichert, T. (2009). Uptake of hydrophilic solutes through plant leaves: current  514  state of knowledge and perspectives of foliar fertilization. Critical Reviews in Plant Sciences  515  28, 36‐68.  516  Fernandez, V., Khayet, M., Montero‐Prado, P., Heredia‐Guerrero, J.A., Liakopoulos, G.,  517  Karabourniotis, G., Del Rio, V., Dominguez, E., Tacchini, I., Nerin, C., Val, J., and Heredia, A.        22    518  (2011). New Insights into the Properties of Pubescent Surfaces: Peach Fruit as a Model.  519  Plant Physiology 156, 2098‐2108.  520  Flechard, C.R., Fowler, D., Sutton, M.A., and Cape, J.N. (1999). A dynamic chemical model of bi‐ 521  directional ammonia exchange between semi‐natural vegetation and the atmosphere.  522  Quarterly Journal of the Royal Meteorological Society 125, 2611‐2641.  523  Fowler, D., Pilegaard, K., Sutton, M.A., Ambus, P., Raivonen, M., Duyzer, J., Simpson, D., Fagerli, H.,  524  Fuzzi, S., Schjoerring, J.K., Granier, C., Neftel, A., Isaksen, I.S.A., Laj, P., Maione, M., Monks,  525  P.S., Burkhardt, J., Daemmgen, U., Neirynck, J., Personne, E., Wichink‐Kruit, R., Butterbach‐ 526  Bahl, K., Flechard, C., Tuovinen, J.P., Coyle, M., Gerosa, G., Loubet, B., Altimir, N.,  527  Gruenhage, L., Ammann, C., Cieslik, S., Paoletti, E., Mikkelsen, T.N., Ro‐Poulsen, H., Cellier,  528  P., Cape, J.N., Horvath, L., Loreto, F., Niinemets, U., Palmer, P.I., Rinne, J., Misztal, P.,  529  Nemitz, E., Nilsson, D., Pryor, S., Gallagher, M.W., Vesala, T., Skiba, U., Brueggemann, N.,  530  Zechmeister‐Boltenstern, S., Williams, J., O'dowd, C., Facchini, M.C., De Leeuw, G.,  531  Flossman, A., Chaumerliac, N., and Erisman, J.W. (2009). Atmospheric composition change:  532  Ecosystems‐Atmosphere interactions. Atmospheric Environment 43, 5193‐5267.  533  Freer‐Smith, P.H., Beckett, K.P., and Taylor, G. (2005). Deposition velocities to Sorbus aria, Acer  534  campestre, Populus deltoides X trichocarpa 'Beaupre', Pinus nigra and X Cupressocyparis  535  leylandii for coarse, fine and ultra‐fine particles in the urban environment. Environmental  536  Pollution 133, 157‐167.  537  Fuentes, J.D., and Gillespie, T.J. (1992). A gas‐exchange system to study the effects of leaf surface  538  wetness on the deposition of ozone. Atmospheric Environment Part a‐General Topics 26,  539  1165‐1173.  540  Gillespie, T.J., and Kidd, G.E. (1978). Sensing duration of leaf moisture retention using electrical‐ impedance grids. Canadian Journal of Plant Science 58, 179‐187.  541  542  Herminghaus, S., Brinkmann, M., and Seemann, R. (2008). Wetting and dewetting of complex  543  surface geometries. Annual Review of Materials Research 38, 101‐121.        23    544  Holloway, P.J. (1969). Chemistry of Leaf Waxes in Relation to Wetting. Journal of the Science of  545  546  Food and Agriculture 20, 124‐128.  Huber, L., and Gillespie, T.J. (1992). Modeling leaf wetness in relation to plant‐disease  547  548  epidemiology. Annual Review of Phytopathology 30, 553‐577.  Hunsche, M., and Noga, G. (2012). Effects of relative humidity and substrate on the spatial  549  association between glyphosate and ethoxylated seed oil adjuvants in the dried deposits of  550  sessile droplets. Pest Management Science 68, 231‐239.  551  IFA, 2012. GESTIS‐database on hazardous substances  http://www.dguv.de/ifa/en/gestis/stoffdb/index.jsp, assesed 8th September, 2013.  552  553  Jacob, D.J. (2000). Heterogeneous chemistry and tropospheric ozone. Atmospheric Environment  554  555  34, 2131‐2159.  Khayet, M., and Fernandez, V. (2012). Estimation of the solubility parameters of model plant  556  surfaces and agrochemicals: a valuable tool for understanding plant surface interactions.  557  Theoretical Biology and Medical Modelling 9.  558  Kumar, A., and Whitesides, G.M (1994) Patterned Condensation Figures as Optical Diffraction 559  560  Gratings Science 263, 60-62.  Lamaud, E., Loubet, B., Irvine, M., Stella, P., Personne, E., and Cellier, P. (2009). Partitioning of  561  ozone deposition over a developed maize crop between stomatal and non‐stomatal  562  uptakes, using eddy‐covariance flux measurements and modelling. Agricultural and Forest  563  Meteorology 149, 1385‐1396.  564  Launiainen, S., Katul, G.G., Gronholm, T., and Vesala, T. (2013). Partitioning ozone fluxes between  565  canopy and forest floor by measurements and a multi‐layer model. Agricultural and Forest  566  Meteorology 173, 85‐99.        24    567  Liao, K.S., Fu, H., Wan, A., Batteas, J.D., and Bergbreiter, D.E. (2009). Designing Surfaces with  568  Wettability That Varies in Response to Solute Identity and Concentration. Langmuir 25, 26‐ 569  28.  570  Lo Nostro, P., and Ninham, B.W. (2012). Hofmeister Phenomena: An Update on Ion Specificity in  571  572  Biology. Chemical Reviews 112, 2286‐2322.  Mauer, L.J., and Taylor, L.S. (2010). "Water‐Solids Interactions: Deliquescence," in Annual Review  573  of Food Science and Technology, Vol 1, eds. M.P. Doyle & T.R. Klaenhammer.  (Palo Alto:  574  Annual Reviews), 41‐63.  575  Monteith, J.L. (1957). Dew. Quarterly Journal of the Royal Meteorological Society 83, 322‐341.  576  Neinhuis, C., and Barthlott, W. (1998). Seasonal changes of leaf surface contamination in beech,  577  oak, and ginkgo in relation to leaf micromorphology and wettability. New Phytologist 138,  578  91‐98.  579  Oum, K.W., Lakin, M.J., Dehaan, D.O., Brauers, T., and Finlayson‐Pitts, B.J. (1998). Formation of  580  molecular chlorine from the photolysis of ozone and aqueous sea‐salt particles. Science  581  279, 74‐77.  582  Pariyar, S., Eichert, T., Goldbach, H.E., Hunsche, M., and Burkhardt, J. (2013). The exclusion of  583  ambient aerosols changes the water relations of sunflower (Helianthus annuus) and bean  584  (Vicia faba) plants. Environmental and Experimental Botany 88, 43‐52.  585  Pegram, L.M., and Record, M.T. (2007). Hofmeister salt effects on surface tension arise from  586  partitioning of anions and cations between bulk water and the air‐water interface. Journal  587  of Physical Chemistry B 111, 5411‐5417.  588  Pilinis, C., Seinfeld, J.H., and Grosjean, D. (1989). Water content of atmospheric aerosols.  589  Atmospheric Environment 23, 1601‐1606.        25    590  Pleijel, H., Karlsson, G.P., Danielsson, H., and Sellden, G. (1995). Surface wetness enhances ozone  591  deposition to a pasture canopy. Atmospheric Environment 29, 3391‐3393.  592  Popek, R., Gawronska, H., Wrochna, M., Gawronski, S.W., and Saebo, A. (2013). Particulate matter  593  on foliage of 13 woody species: deposition on surfaes and phytostabilisation in waxes ‐ a 3‐ 594  year study. International Journal of Phytoremediation 15, 245‐256.  595  Pöschl, U. (2005). Atmospheric aerosols: Composition, transformation, climate and health effects.  596  597  Angewandte Chemie‐International Edition 44, 7520‐7540.  Roeselova, M., Jungwirth, P., Tobias, D.J., and Gerber, R.B. (2003). Impact, trapping, and  598  accommodation of hydroxyl radical and ozone at aqueous salt aerosol surfaces. A  599  molecular dynamics study. Journal of Physical Chemistry B 107, 12690‐12699.  600  Rosado, B.H.P., and Holder, C.D. (2013). The significance of leaf water repellency in ecohydrological  601  research: a review. Ecohydrology 6, 150‐161.  602  603  604  Rosenfeld, D., Lohmann, U., Raga, G.B., O'dowd, C.D., Kulmala, M., Fuzzi, S., Reissell, A., and Andreae,  M.O. (2008). Flood or drought: How do aerosols affect precipitation? Science 321, 1309‐ 1313.  605  Roth‐Nebelsick, A. (2007). Computer‐based studies of diffusion through stomata of different  architecture. Annals of Botany 100, 23‐32.  606  607  Saebo, A., Popek, R., Nawrot, B., Hanslin, H.M., Gawronska, H., and Gawronski, S.W. (2012). Plant  608  species differences in particulate matter accumulation on leaf surfaces. Science of the Total  609  Environment 427, 347‐354.  610  Schindelholz, E., and Kelly, R.G. (2012). Wetting phenomena and time of wetness in atmospheric  611  612  corrosion: a review. Corrosion Reviews 30, 135‐170.  Schönherr, J. (2001). Cuticular penetration of calcium salts: effects of humidity, anions, and  613  adjuvants. Journal of Plant Nutrition and Soil Science‐Zeitschrift für Pflanzenernährung und  614  Bodenkunde 164, 225‐231.        26    615  Schönherr, J., and Bukovac, M.J. (1972). Penetration of stomata by liquids ‐ dependence on surface  616  tension, wettability, and stomatal morphology. Plant Physiology 49, 813‐819.  617  Schuepp, P.H. (1993). Leaf boundary layers. New Phytologist 125, 477‐507.  618  Sentelhas, P.C., Gillespie, T.J., and Santos, E.A. (2007). Leaf wetness duration measurement:  619  comparison of cylindrical and flat plate sensors under different field conditons.  620  International Journal of Biometeorology 51, 265‐273.  621  Stevens, P.J.G. (1993). Organosilicone surfactants as adjuvants for agrochemicals. Pesticide Science  622  623  38, 103‐122.  Sutton, M.A., Nemitz, E., Milford, C., Campbell, C., Erisman, J.W., Hensen, A., Cellier, P., David, M.,  624  Loubet, B., Personne, E., Schjoerring, J.K., Mattsson, M., Dorsey, J.R., Gallagher, M.W.,  625  Horvath, L., Weidinger, T., Meszaros, R., Dammgen, U., Neftel, A., Herrmann, B., Lehman,  626  B.E., Flechard, C., and Burkhardt, J. (2009). Dynamics of ammonia exchange with cut  627  grassland: synthesis of results and conclusions of the GRAMINAE Integrated Experiment.  628  Biogeosciences 6, 2907‐2934.  629  Tang, I.N., Wong, W.T., and Munkelwitz, H.R. (1981). The relative importance of atmospheric  630  sulfates and nitrates in visibility reduction. Atmospheric Environment 15, 2463‐2471.  631  van Hove, L.W.A., and Adema, E.H. (1996). The effective thickness of water films on leaves.  632  633  Atmospheric Environment 30, 2933‐2936.  van Hove, L.W.A., Adema, E.H., Vredenberg, W.J., and Pieters, G.A. (1989). A study of the  634  adsorption of NH3 and SO2 on leaf surfaces. Atmospheric Environment 23, 1479‐1486.  635  Vorholt, J.A. (2012). Microbial life in the phyllosphere. Nature Reviews Microbiology 10, 828‐840.  636  Walker, J.T., Jones, M.R., Bash, J.O., Myles, L., Meyers, T., Schwede, D., Herrick, J., Nemitz, E., and  637  Robarge, W. (2013). Processes of ammonia air‐surface exchange in a fertilized Zea mays  638  canopy. Biogeosciences 10, 981‐998.        27    639  Wichink Kruit, R.J., Jacob, A.F.G., and Holtslaga, A.a.M. (2008). Measurements and estimates of  640  leaf wetness over agricultural grassland for dry deposition modeling of trace gases.  641  Atmospheric Environment 42, 5304‐5316.  642  Wyers, G.P., and Erisman, J.W. (1998). Ammonia exchange over coniferous forest. Atmospheric  Environment 32, 441‐451.  643  644  Xu, L., Zhu, H., Ozkan, H.E., Bagley, W.E., Derksen, R.C., and Krause, C.R. (2010). Adjuvant effects  645  on evaporation time and wetted area of droplets on waxy leaves. Transactions of the  646  ASABE 53, 13‐20.  647  Zhang, Y.J., and Cremer, P.S. (2010). "Chemistry of Hofmeister Anions and Osmolytes," in Annual  648  Review of Physical Chemistry, Vol 61, eds. S.R. Leone, P.S. Cremer, J.T. Groves, M.A.  649  Johnson & G. Richmond.  (Palo Alto: Annual Reviews), 63‐83.  650  651  652  653  654  655  656  657  658  659  660        28    661  Figure 662  663  Figure 1: 664  Measurement of leaf wetness on a potato field, comparing an artificial leaf Campbell Leaf wetness 665  sensor 237 (blue line, upper image; Campbell Scientific, Logan, UT, USA), and a leaf wetness sensor 666  directly attached to a potato leaf (red line, upper image; construction see Burkhardt & Gerchau, 1994) 667  Ambient air humidity (black line, lower picture) and photosyntheticcally active radiation (pink line, 668  lower picture) are also shown CET: Central European Time 669  670  671  Movies 672  Movie 1: 673  Potassium iodide (KI) crystals on a Pinus sylvestris needle under changing humidity in an 674  environmental scanning electron microscope Three deliquescence/efflorescence cycles are shown, 675  cycling between approximately 55% and 70% RH 676  677  Movie 2: 678  Potassium thyocyanate (KSCN) crystals on a Pinus sylvestris needle under increasing humidity in an 679  environmental scanning electron microscope Humidity increases from 60 to 65 % RH 680  681          Figure 1.TIFF Copyright of Frontiers in Plant Science is the property of Frontiers Media S.A and its content may not be copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written permission However, users may print, download, or email articles for individual use ...1    1  ? ?Breath figures? ?? on leaf surfaces – formation and effects of 2  microscopic leaf wetness 3  Juergen Burkhardt1 & M Hunsche2 4  5  6  Revision 7  8  9  10  11  University of Bonn, Karlrobert-Kreiten-Str... Depending on the amount and concentration of 31  the dissolved ions, the physicochemical properties of microscopic leaf wetness can be considerably 32  different from those of pure water Microscopic leaf. .. exchange of ions Microscopic leaf wetness can also enhance the 35  dissolution, the emission, and the reaction of specific atmospheric trace gases e.g ammonia, SO2, or 36  ozone, leading to a strong

Ngày đăng: 01/11/2022, 08:59

Xem thêm:

TÀI LIỆU CÙNG NGƯỜI DÙNG

TÀI LIỆU LIÊN QUAN