COBISS: 1.02 CONSIDER A CYLINDRICAL CAVE: A PHYSICIST'S VIEW OF CAVE AND KARST SCIENCE VZEMIMO VALJASTO JAMO: POGLED FIZIKA NA ZNANOST O JAMAH IN KRASU Matthew D. COVINGTON1 & Matija PERNE2 Abstract UDC 551.44:53 Matthew D. Covington & Matija Perne: Consider a cylindrical cave: A physicist's view of cave and karst science We review the current understanding of the physics of caves and karst. Our review focuses on research that has used simple physically based models to improve understanding of processes that occur in karst. The topics we cover include cave atmosphere dynamics, transport within karst conduits, and models of speleogenesis and related processes. We highlight recent advances in these subjects and attempt to identify promising areas for future work. In our judgment, many of the most intriguing open questions relate to the interactions between these three groups of processes. Keywords: Karst, speleology, physics, mathematical modeling, cave meteorology, hydrology, speleogenesis. Izvleček UDK 551.44:53 Matthew D. Covington & Matija Perne: Vzemimo valjasto jamo: pogled fizika na znanost o jamah in krasu V članku pregledava trenutno poznavanje fizike jam in krasa. Pri tem se osredotočava na raziskave, ki so razumevanje kraških procesov poglobile z uporabo preprostih modelov na osnovi fizike. Obravnavava vedenje jamskega ozračja, transport v kraških kanalih in modele nastanka jam ter povezanih procesov. Izpostavljava sodobna dognanja na teh področjih in iščeva obetavne teme za nadaljnje raziskave. Po najinem mnenju so mnoga med bolj privlačnimi odprtimi vprašanji povezana z medsebojnim vplivom med obravnavanimi tremi skupinami procesov. Ključne besede: Kras, speleologija, fizika metematično modeliranje, jamska meteorologija, speleogeneza. INTRODUCTION When a colleague excitedly showed Eugene Wigner the result of a complex quantum mechanical calculation produced by a computer, Wigner's storied reply was, "It is nice to know that the computer understands the problem, but I would like to understand it too (Heller & Tomsovic 1993)." This reply reflects a general attitude in theoretical physics, that one has not really understood something until one has an analytical mathematical model for it. While computers play an increasingly dominant role in quantitative science, and we are more and more awash with data, analytical models retain an important function. Often the results of computer simulations can be difficult to generalize beyond the particular cases run. Analytical models can provide a powerful tool for understanding the results of these simulations and illuminating relevant general principles. They can play a very similar role in data analysis. Within the field of physics, there is arguably a bias toward the analytical, the simple, the elegant. However, it is certainly a bias that has served physics well (Wigner 1960), along with many other fields. 1 University of Arkansas Department of Geosciences Fayetteville, Arkansas, USA, e-mail: mcoving@uark.edu 2 University of Arkansas Department of Geosciences Fayetteville, Arkansas, USA, Jožef Stefan Institute, Department of Systems and Control Ljubljana, Slovenia Received/Prejeto: 02.03.2015 ACTA CARSOLOGICA 44/3, 363-380, POSTOJNA 2015 MATTHEW D. COVINGTON & MATIJA PERNE It is our task in this article to review the physics of caves. Given that basic physics underpins our understanding of a wide variety of processes that occur in caves and karst, we must choose a narrower lens through which to view the topic. The lens that we have chosen is that of the simple physics-based model. There is an increasingly well-worn path into karst science that has been trodden by physicists. Most of these scientists have entered karst science as physicist cavers, whose passion and curiosity about the underground world inspired their scientific contributions to karst studies (e.g. the interview of Wolfgang Dreybrodt in Lucic 2011). The work done by this group of physicists has often focused on simple and general models. This work has employed analytical solutions, dimensional analysis, and simple numerical models to enable understanding of more complex experimental and observational work. Therefore, in choosing to focus on simple models, we have also chosen to focus on the type of work that physicists have most often undertaken when they have delved into the realm of karst. We also focus more heavily on recent contributions, in hopes of illuminating promising areas for future work. CAVE CLIMATE AND METEOROLOGY The study of cave atmospheres has frequently attracted researchers with a background in physics. Perhaps this results from the ease with which the laws of physics can be applied to the problem, or perhaps from the curiosity of cavers who are always following the wind. The two most complete works on cave atmospheres have been written from a physics perspective (Badino 1995; Lismonde 2002), and a prior review of cave physics devoted about half of its space to this topic (Wigley & Brown 1976). Cave atmospheres are known for their constancy in comparison to the surface atmosphere. However, cave atmospheres are not truly constant, and it is their variability in space and time that poses many of the most interesting questions and most relevant unknowns. The physics of cave atmospheres was recently reviewed by Badino (2010), who divides the field into "cave climatology," the study of the average cave atmospheric conditions that vary slowly in time, and "cave meteorol- ogy," the study of how the cave fluctuates around this average condition over relatively short timescales. We adopt this division here, as it seems an apt analogy to the traditional fields of climatology and meteorology. However, there is a difference in scale and degree. While a meteorologist often studies relatively dramatic phenomena, a cave meteorologist may study diurnal or seasonal variations on the order of 0.1 °C and humidity variations of a few percent. Understanding cave atmospheres, their variability, and the factors that control them is increasingly important as we seek to interpret paleoclimate records from caves (Fairchild et al. 2006). The dynamics of cave atmospheres also has important implications for cave ecosystems (e.g. Culver 2005; Tobin et al. 2013), the protection of caves from anthropogenic impacts (e.g. Cigna 1993; Hoyos et al. 1998), and the formation and evolution of caves over time (e.g. Dreybrodt et al. 2005b; Covington et al. 2013). ! Cathy, Switzerland Temperate Muruk, New Guinea Eauatorial wet Kijahe Xontjoa, Mexico Tropical wet 2.5 3.0 3.54.04.5 5.0 5.5 Temperature (°C) Kievskaya, Uzbekistan Q Semi-arid 13 14 15 16 17 18 Temperature (°C) 10 11 12 13 14 15 16 Temperature (°C) 3.54.04.5 5.0 5.5 6.0 6.5 Temperature (°C) Fig. 1: Temperature profiles with depth in deep cave systems in different climatic settings. Reproduced using data from Luetscher & Jeannin (2004). 364 ACTA CARSOLOGICA 44/3- 2015 CONSIDER A CYLINDRICAL CAVE: A PHYSICIST'S VIEW OF CAVE AND KARST SCIENCE LARGE SCALE THERMAL DYNAMICS OF KARST AQUIFERS The average local temperature on the surface exerts a first-order control on cave temperature. Therefore cave temperature is strongly dependent on both altitude and latitude. More specifically, the temperature is primarily controlled by the average temperature of the fluids that flow through the aquifer, both air and water (Luetscher & Jeannin 2004; Badino 2010). Karst aquifers receive geothermal flux from below, and a heat flux from above that is driven by surface temperature. However, for the unsaturated zone of unconfined karst aquifers, the geo-thermal and surface heat flow rates are typically dwarfed by the heat capacity rate of the fluids that cross the aquifer, such that the temperature inside the aquifer is approximately equilibrated to the average temperature at the surface at the same altitude (Bogli 1980; Luetscher & Jeannin 2004; Badino 2005). Karst can be considered an end-member case among aquifers, where advective heat transport dominates over conductive processes. This can be expressed quantitatively by stating that Peclet numbers for heat transport are large within karst systems (Domenico & Palciauskas 1973), where the Peclet Number is a ratio of the advective and conductive heat transport rates. Consequently, in deep unsaturated zones, once below the shallow surface-influenced zone, karst aquifers display a systematic increase in temperature with depth that is typically much less than the normal geothermal gradient of approximately 2.5 °C/100 m. Observed thermal gradients in deep caves (Fig. 1) are between the values of the energy dissipation rate of falling water (0.234 °C/100 m), and the adiabatic lapse rate of moist air (0.5 °C/100 m) (Luetscher & Jeannin 2004). Luetscher & Jeannin (2004) argue from estimates of air flux in two caves (Holloch and La Diau) that the energy flux due to air circulation is 2 to 20 times larger than the energy flux due to water. They cite as further evidence that many of the observed caves display thermal gradients close to the adiabatic lapse rate of moist air. However, Badino (2010) asserts that these authors overestimate typical air flux and concludes that water is the dominant factor in most settings. In either case, observed temperature gradients typically lie between those expected by the dominance of air and water. Climate also appears to be an important factor in determining temperature profiles, with caves in wetter climates displaying lower gradients (i.e. more water dominated) than in drier climates (Fig. 1). Many of the temperature profiles also display reduced gradients within the deeper portion of the cave, where the influence of air is reduced. The debate on the relative importance of water and air in determining thermal profiles highlights a need for further work to constrain the flux of air through karst systems. The thermal response of a karst massif to change in climate has also been considered using simple models (Badino 2004). The temperature of a karst massif is roughly equal to the average temperature of the fluids that cross it. However, if climate is changing, then the temperature of these fluids may also change with time. Since the karst massif has a large heat capacity, this change will not be instantaneous and will occur over some timescale. Badino (2004) suggests that a timescale of particular interest is heat capacity timescale, which is the time over which the heat capacity of the fluids crossing the massif is equal to the heat capacity of the rock within the massif. This can be written as T = IhPAK (1) Tcap \Cfp{jR, (1) where cr and cf are the specific heat capacities of the rock and fluid (water or air), pr and pf are the densities of the rock and fluid, H is the thickness of the massif, and R (dimension of L/T) is the flux of water or air. In the case of water, annual recharge can be used for R. The ratio in parentheses in Equation 1 is roughly equal to 0.5 for water and 1500 for air. Considering recharge by water at a rate of 1 m yr-1 would lead to a heat capacity timescale of 50 years for a rock thickness of H = 100 m and 500 years for a thickness of H = 1000 m. These values would suggest that the massif would lag behind local climate changes by the order of a few hundred years. However, there are other potentially relevant timescales. In particular, as also noted by Badino (2004), a temperature pulse will propagate into a rock body via conduction to a depth H over a timescale given by Tcond ~ H2/ ar, (2) where ar is the thermal diffusivity of rock (~10-6 ms-1 for dry rock). In order for the entire massif to change temperature there are two requirements: 1) the heat capacity of the fluids that have crossed it has to be comparable to or greater than the heat capacity of the massif, and 2) the temperature must have time to conduct away from areas of fluid contact and through the body of the rock. Therefore, if the conduction timescale is much longer than the heat capacity timescale, it would suggest an influence of conduction on the response time of the karst massif. In fact, Equation 2 implies quite long timescales for the equilibration of large thicknesses of rock. For example H = 1000 m would lead to an equilibration time scale of Tcond ~3 x 104 yr. However, because of the network of conduits that penetrate the aquifer, it is unlikely that heat within a karst aquifer will need to conduct through its ACTA CARSOLOGICA 44/3 - 2015 365 MATTHEW D. COVINGTON & MATIJA PERNE Fig. 2: Air temperature in Blowing Springs Cave, Arkansas, USA, as a function of distance into the cave. Later in the winter, cold outside air penetrates deeper than in the late fall. Temperature profiles are shown as the difference between cave temperature and external temperature normalized by the difference between equilibrium cave temperature and the external temperature. This shows that cooling is not simply a result of cooler outside temperature but rather an increase in the thermal penetration length. entire thickness. Therefore, half of typical distance between large conduits may be a more appropriate value for H than the entire aquifer thickness. These two timescales assume a decoupling between conduction and heat exchange due to fluid flow. Equation 1 makes an assumption that the fluids are able to exchange all available heat, whereas Equation 2 assumes that the temperature at the fluid rock boundary is coupled to the surface temperature. The processes of fluid heat exchange and conduction are actually coupled, and their coupling leads to a third relevant timescale, which is the timescale over which a thermal pulse can propagate a given distance, L, down a conduit imbedded in rock, 16arL2 coupled 7t4/2_D 2 V2 ' (3) where ^ = (pf cp,f )/(prc ) is the ratio of the densities and specific heat capacities of the fluid and rock, DH is the conduit hydraulic diameter, and V is the fluid flow velocity. This can be derived from the thermal length scale given in Equation 22 of Covington et al. (2012b). In general, thermal pulses do not move down conduits at the same velocity as the fluid. This results because of exchange of heat between the fluid and rock. The pulse is damped as it flows along the conduit, but over time the rock cools or heats and the thermal pulse propagates further. This pulse propagation timescale may be the most important one to determine the long-term temperature behavior of rock immediately surrounding conduits and the fluids within the caves themselves. Though an overall picture has emerged, a variety of questions remain unexplored regarding the importance of these different timescales, and the internal aquifer structure, in determining the long-term thermal behavior of karst aquifers. HEAT EXCHANGE WITHIN KARST CONDUITS Covington et al. (2011) explored the relative importance of mechanisms of heat exchange in karst conduits, as there were inconsistencies between prior models of karst conduit heat exchange. Some models assumed that heat exchange was limited by convective exchange in the boundary layer near the wall (Wigley & Brown 1971; Long & Gilcrease 2009), other models assumed that heat conduction within the wall was limiting (Benderitter et al. 1993), and others accounted for both processes (Liedl & Sauter 1998; Birk et al. 2006). Covington et al. (2011) showed that the relative importance of convective and conductive heat exchange is determined by a critical time scale m k2«,Nu2' (4) where kr and kw are the thermal conductivities of rock and water, respectively, DH is the hydraulic diameter of the conduit, a is the thermal diffusivity of rock, and Nu is the Nusselt number. For temperature pulses with 364 ACTA CARSOLOGICA 44/3- 2015 t CONSIDER A CYLINDRICAL CAVE: A PHYSICIST'S VIEW OF CAVE AND KARST SCIENCE Fig. 3: Temperature time series from near Camp 3 in Sistema J2 at a depth of approximately -1100 m demonstrate a complex variability with time throughout the wet season. timescales tpulse ^eq (7) where L is the conduit length, Vh is the hydraulic gradient, a0 is the initial aperture, kn2 is the kinetic rate constant for non-linear calcite dissolution near equilibrium (dimension of L-2T-1N), and Ceq is the equilibrium concentration of calcite (dimension of L-3N). C is a constant that depends on the shape of the conduit and is approximately equal to 6.1 x 10-3 m-5/3 mol s4/3 for square and circular cross sections and 6.1 x 10-4 m-5/3 mol s4/3 for fracturelike cross sections. The scalings seen in Equation 7, were also reproduced with an analytical approximation (Dreybrodt 1996; Dreybrodt & Gabrovsek 2000) and arguably provide us with the deepest understanding that we currently have about the timescale of karstification and the factors that control it. More complex dynamics arise as one moves from 1D fractures to two-dimensional (2D) representations of fractures or 2D networks of fractures. Hanna & Raja-ram (1998) showed that aperture heterogeneity within a fracture can result in the formation of preferential flow paths that accelerate breakthrough in comparison to the 1D case. Similarly, exchange flows between fractures and matrix, or larger and smaller aperture fractures within a network, can also accelerate breakthrough (Bauer et al. 2003; Gabrovsek et al. 2004). Szymczak & Ladd (2011) demonstrate that the propagation of a dissolution front within a fracture is fundamentally unstable, which results in fingering of the dissolution front. The instability accelerates breakthrough, but a newer formulation 364 ACTA CARSOLOGICA 44/3- 2015 CONSIDER A CYLINDRICAL CAVE: A PHYSICIST'S VIEW OF CAVE AND KARST SCIENCE of breakthrough time that accounts for these effects remains elusive (Szymczak & Ladd 2012). Simulations of the evolution of 2D fracture networks have enabled studies of the evolution of cave plan forms (Groves & Howard 1994; Siemers & Dreybrodt 1998) and cave profiles (Gabrovsek & Dreybrodt 2001). Such models have been used to explore the competition between different flow paths and the influence of mixing corrosion (Gabrovsek & Dreybrodt 2000), the effect of CO2 sources (Gabrovsek et al. 2000), the formation of flank margin caves (Dreybrodt & Romanov 2007; Dreybrodt et al. 2009), buoyant convection (Chaudhuri et al. 2009), and karstification around dam sites (Drey-brodt et al. 2002). Double-porosity models, where flow through discrete conduits is coupled to the flow through porous rock matrix, have also been developed (Kaufmann & Braun 2000; Liedl et al. 2003). Kaufmann (2009) introduced a three-dimensional karst evolution model that coupled speleogenesis and landscape evolution. For a comprehensive review of fracture network speleogen-esis models, which also presents some novel results, see Dreybrodt et al. (2005a). THE NEXT GENERATION OF SPELEOGENESIS MODELS Speleogenetic models have primarily focused on the early stages of cave formation and the dynamics of flow network initiation. However, there is a rich host of processes that occur in the later stages of speleogenesis that have received little modeling attention. We have only recently seen the first network speleogenetic models that consider the transition to open channel flow and its potential role in preferential selection of flow paths (Perne et al. 2014b). Mature cave systems often develop undercapture routes, though this will only happen if lower routes are able to enlarge quickly enough to outpace the downcutting of the active stream passage. Gabrovsek et al. (2014) derive a dimensionless number, called the Loop-to-Canyon-Ratio, that is the ratio of the timescales for breakthrough of the lower passage and downcutting of the active stream passage. They use this ratio to explore the controls on multi-level cave development and cave evolution within the epiphreatic zone. Turbulent flow dominates the later stages of cave formation. There are unresolved questions concerning dissolution rates under turbulent conditions (Hammer et al. 2011; Covington 2014). Direct application of the theory would suggest that surface reaction rates are limiting under turbulent flow conditions. However, scallops and flutes are features that strongly suggest that dissolution rates are a function of flow structure (Blumberg & Curl 1974). It may be that chemomechanical processes play an important role, whereby individual grains are chemically loosened and then mechanically plucked. High resolution scanning of dissolving surfaces suggests that grain detachment may strongly influence rates of erosion (Emmanuel & Levenson 2014). Whether or not chemo-mechanical erosion processes are important, mechanical erosion is certain to be important in more powerful cave streams (Newson 1971). However, little is known about controls on the relative importance of chemical and mechanical erosion processes in cave streams, and models have not yet included mechanical processes. Mechanical erosion processes typically scale with the shear stress to a power of 1 to 3 (Whipple et al. 2000). In contrast, transport limited dissolution scales with shear stress to the 1/3 to 1/2 power (Opdyke et al. 1987). The controls on the variability of dissolution rates in cave streams are not well understood, but preliminary work suggests that chemically driven changes in dissolution rates within surface streams tend to scale weakly with discharge (Covington et al. 2015). There is a broad push within the geomorphology community to develop mechanistic models of earth surface processes (Dietrich et al. 2003). Mechanistic models for erosion by bedload, abrasion, and plucking (Sklar & Dietrich 2004; Chatanantavet & Parker 2009; Lamb et al. 2008) may prove useful within the next generation of speleogenesis models. Prescriptions for sediment dynamics will also be required to simulate the later stages of cave evolution (Farrant & Smart 2011). There is substantial interest in quantifying the controls on bedrock channel widths, as width is one of the least understood degrees of freedom available to accommodate channel response to contrasts in rock properties, uplift, and climate (e.g. Montgomery & Gran 2001; Finnegan et al. 2005; Yanites & Tucker 2010). Cave channels provide an interesting environment to examine such questions. Records of channel evolution are often well-preserved within caves, and many conceptual models have been developed to understand different cave passage cross sectional shapes (Lauritzen & Lundberg 2000). The cross sections of fossil cave passages may provide clues to past hydrological or climatic conditions. Additionally, due to the absence of hillslopes, the dynamics of cave channel width may be somewhat simpler than surface channels. The records of channel evolution that are preserved underground may prove useful to constrain models of bedrock channel width more broadly. Speleogenesis models that incorporate cross-section evolution have only begun to be developed (Perne 2012; Perne et al. 2014a; Cooper et al. 2014, Fig. 6). The formation and evolution of hypogene cave systems has seen increased attention in the recent past. However, little work has been done to quantitatively model such systems. Birk et al. (2005) examined the ACTA CARSOLOGICA 44/3 - 2015 365 MATTHEW D. COVINGTON & MATIJA PERNE Fig. 6: Initial results from new models of cave channel cross section evolution that use calculations of boundary shear stress along the wall to evolve the channel. (a) A model that uses computational fluid dynamics to calculate shear stress (Perne et al. 2014a). Blue depicts air, and red depicts water. (b) A simpler model that approximates boundary shear stress along a conduit wall with an irregular shape (Cooper et al. 2014). The gray line shows boundary shear stress for the inset conduit cross-section with scale. development of gypsum maze caves in an artesian setting, and a series of studies has examined dissolution under cooling and buoyantly driven flows (Andre & Ra-jaram 2005; Chaudhuri et al. 2008, 2013). Little mathematical modeling work has been done on sulphuric acid speleogenesis. There is substantial debate in the karst community concerning hypothesized diagnostic features of hypogene speleogenesis, such as the morphologic suite of rising flow (Klimchouk 2007), and whether these features must form via deep rising flow or whether other processes such as condensation corrosion, freshwater/saltwater mixing, paragenesis, and flood water might produce similar features (Curl 1966; Mylroie 2008; Audra et al. 2009; Stafford et al. 2009; Palmer 2011). While conceptual models exist for the formation of these features, the proposed mechanisms have not generally been studied using mathematical models. Therefore the plausibility of the various mechanisms is uncertain from a physics perspective, and this area seems ripe for study using more mechanistically based models. Another area of research where substantial advances are likely is the interaction between cave atmospheres and speleogenetic processes. Cave meteorology, in particular air flows, can influence the aggressivity of the water flowing through caves via exchange of CO2 between air and water (Covington et al. 2013). These effects have not yet been included in speleogenetic models. Coupled models of CO2 within cave air and water may first require further observational studies of cave streams and atmospheres to better quantify the controls on CO2 concentrations and their variability (Milanolo & Gabrovšek 2015; Baldini 2010). Meteorology also affects the formation of caves through condensation. The amount of condensed water can be significant (Dublyansky & Dublyansky 2000), and as it initially contains dissolved carbon dioxide but no minerals it is typically fairly aggressive (Dreybrodt et al. 2005b). Condensation corrosion has been proposed to explain the formation of large cupolas (Audra et al. 2002). Condensation on cave walls occurs either continuously, in steady state, or periodically, as a result of temperature variations. Steady state condensation requires a source of water that is warmer than the surroundings (Sarbu & Lascu 1997), and its rate is limited by heat conduction through the bulk of the rock away from the cave wall. The geometry of the cave and the surrounding rock has a strong influence on the rate (Dreybrodt et al. 2005b). In the case of periodic condensation, temperature variations cause heat to be stored in a layer of rock surrounding the cave and dispersed back during colder periods. During the periods when the air is sufficiently warm and moist and heat is being stored, condensation occurs. The total amount of condensation depends on the amplitude and frequency of the temperature signal, and the rock layer thickness required for heat storage is smaller for higher frequencies of temperature variations. Strong daily variations can, for example, cause significant condensation and corrosion even on speleothems (Tarhule- 364 ACTA CARSOLOGICA 44/3- 2015 CONSIDER A CYLINDRICAL CAVE: A PHYSICIST'S VIEW OF CAVE AND KARST SCIENCE Lips & Ford 1998). However, when the total amount of condensation per cycle is small, the water may not drip away but evaporate back during the drying period and re-precipitate the dissolved minerals. In this way, weathered rinds can form (Auler & Smart 2004). Physically based models of the growth of deposi-tional forms within caves have also been developed. Stalactite shape was modeled, and a simple general shape that fits many real stalactites was found (Short et al. 2005). Shapes of stalagmites forming in either steady-state or variable conditions were explained through numerical modeling as well (Romanov et al. 2008). The development of crenulations on speleothems was studied through a stability analysis that demonstrated that the migration pattern of these forms within a speleothem is correlated to film flow rates (Camporeale & Ridolfi 2012). Speleothems are useful for reconstructing paleo-climate (Harmon et al. 1978; Baker et al. 1993), and numerical models of their formation are being used in this context (Mariethoz et al. 2012). Relatively few physics-based models have been developed for karst surface processes or for processes in the epikarst and vadose zone. Gabrovsek (2007) developed a simple model for the vertical distribution of dissolution in a karst aquifer. Using the characteristic length scale for dissolution in vertical fractures, Gabrovsek (2007) examined the assumptions behind the maximum denudation models that use recharge and equilibrium calcium concentrations to estimate denudation rates in a karst terrain. He finds that the maximum denudation formulation is reasonable in most cases, even though not all dissolution occurs at the surface. A number of studies have shown that CO2 concentrations can increase substantially with depth in the vadose zone (e.g. Atkinson 1977; Wood 1985), and recent work suggests that high levels of CO2 may be primarily responsible for dissolution in eo-genetic karst settings rather than mixing corrosion (Gul-ley et al. 2014, 2015). Additionally, Covington (in press) uses dimensional analysis of models of CO2 transport in the vadose zone to suggest that advection of both air and water are important processes in determining the spatial and temporal distributions of CO2. Vertical changes in the partial pressure of CO2 within karst systems have not typically been considered in karst evolution models, and these may be important in determining the distribution of dissolution rates throughout the system. CONCLUSIONS Scientific research often benefits from the interaction between disparate fields. There is a long and continuing history of physicists working within the field of cave and karst science. We argue that this work has provided a substantial contribution to the field, largely as a result of a difference in approach. The physicist is driven to find general mathematical descriptions for the behavior of a system. When dealing with complex systems, a common approach within physics is to develop relatively simple models, sometimes called "toy models," that capture the essence of the dynamics. When successful, this approach provides a powerful tool for understanding and gener- alization. It can aid in the interpretation of numerical simulations, experiments, and observational data. Simple models have been and continue to be applied to processes within karst. They have provided a general framework for understanding a variety of phenomena, from cave climate and meteorology, to karst transport, to speleogen-esis. This work is hardly done, and there are many open questions that we have attempted to elucidate above. In our judgement, many of the most exciting potential advances relate to the interactions between these three sets of processes. ACKNOWLEDGMENTS We thank Wolfgang Dreybrodt, whose headlamp has shown the way, and Franci Gabrovsek and Lee Florea for helpful reviews that substantially improved this work. We also acknowledge inspiration for the title from John Harte's book "Consider a Spherical Cow" (Harte 1998). M.D.C. and M.P. acknowledge support from the National Science Foundation under grant no. 1226903, M.P. acknowledges support of the Slovenian Research Agency through Research Programme P2-0001. The data displayed in Fig. 2 were collected by students in M.D.C.'s fall 2014 Karst Hydrogeology course. ACTA CARSOLOGICA 44/3 - 2015 365 MATTHEW D. COVINGTON & MATIJA PERNE REFERENCES Andre, B.J. & H. Rajaram, 2005: Dissolution of limestone fractures by cooling waters: Early development of hypogene karst systems.- Water Resources Research, 41, 1-16. Ashton, K., 1966: The analysis of flow data from karst drainage systems.- Transactions of the Cave Research Group of Great Britain, 7, 2, 163-203. Atkinson, T., 1977: Carbon dioxide in the atmosphere of the unsaturated zone: An important control of groundwater hardness in limestones.- Journal of Hydrology, 35, 1-2, 111-123. Audra, P., Bigot, J.Y. & L. Mocochain, 2002: Hypogenic caves in Provence (France): Specific features and sediments.- Acta Carsologica, 31, 3, 33-50. Audra, P., Mocochain, L., Bigot, J.Y. & J.C. Nobecourt, 2009: Hypogene cave patterns. In: Hypogene Spe-leogenesis and Karst Hydrogeology of Artesian Basins. Special Paper, 1, Ukrainian Institute of Speleology and Karstology. Auler, A.S. & P.L. Smart, 2004: Rates of condensation corrosion in speleothems.- Speleogenesis and evolution of karst aquifers, 2, 2. Badino, G., 1995: Fisica del Clima Sotterraneo-Memorie IIS.- Volume 7. Istituto Italiano di Speleologia, Bologna, Italy. Badino, G., 2004: Cave temperatures and global climate change.- International Journal of Speleology, 33, 1, 103-114. Badino, G., 2005: Underground drainage systems and geothermal flux.- Acta Carsologica, 34, 2, 277-316. Badino, G., 2010: Underground meteorology - What's the weather underground? Podzemna meteorologija: Kakšno je vreme v podzemlju?.- Acta Carsologica, 39, 3, 427-448. Baker, A., Smart, P.L., Edwards, R.L. & D.A. Richards, 1993: Annual growth banding in a cave stalagmite.-Nature, 364, 518-520. Baldini, J.U.L., 2010: Cave atmosphere controls on stalagmite growth rate and palaeoclimate records.-Geological Society, London, Special Publications, 336, 283-294. Banner, J.L., Guilfoyle, A., James, E.W., Stern, L.A. & M. Musgrove, 2007: Seasonal Variations in Modern Speleothem Calcite Growth in Central Texas, U.S.A..- Journal of Sedimentary Research, 77, 1994, 615-622. Bauer, S., Liedl, R. & M. Sauter, 2003: Modeling of karst aquifer genesis: Influence of exchange flow.- Water Resources Research, 39, 10, 1-12. Benderitter, Y., Roy, B. & A. Tabbagh, 1993: Flow characterization through heat transfer evidence in a carbonate fractured medium: first approach.- Water resources research, 29, 11, 3741-3747. Birk, S., Liedl, R. & M. Sauter, 2006: Karst spring responses examined by process-based modeling.-Groundwater, 44, 6, 832-836. Birk, S., Liedl, R., Sauter, M. & G. Teutsch, 2005: Simulation of the development of gypsum maze caves.-Environmental Geology, 48, 3, 296-306. Blumberg, P. & R.L. Curl, 1974: Experimental and theoretical studies of dissolution roughness.Journal of Fluid Mechanics, 65, 4, 735-751. Bogli, A., 1980: Karst Hydrology and Physical Speleology.- Springer-Verlag. Breecker, D.O., Payne, A.E., Quade, J., Banner, J.L., Ball, C.E., Meyer, K.W. & B.D. Cowan, 2012: The sources and sinks of CO2 in caves under mixed woodland and grassland vegetation.- Geochimica et Cos-mochimica Acta, 96, 230-246. Buecher, R.H., 1999: Microclimate study of Kartchner Caverns, Arizona.- Journal of Cave and Karst Studies, 61, 108-120. Buhmann, D. & W. Dreybrodt, 1985a: The kinetics of calcite dissolution and precipitation in geologically relevant situations of karst areas: 1. Open system.-Chemical Geology, 48, 1-4, 189-211. Buhmann, D. & W. Dreybrodt, 1985b: The kinetics of calcite dissolution and precipitation in geologically relevant situations of karst areas: 2. Closed system.-Chemical Geology, 53, 1-2, 109-124. Camporeale, C. & L. Ridolfi, 2012: Hydrodynamic-driven stability analysis of morphological patterns on stalactites and implications for cave paleoflow reconstructions.- Physical review letters, 108, 23, 238501. Chatanantavet, P. & G. Parker, 2009: Physically based modeling of bedrock incision by abrasion, plucking, and macroabrasion.- Journal of Geophysical Research, 114, F4, 1-22. Chaudhuri, A., Rajaram, H. & H. Viswanathan, 2008: Alteration of fractures by precipitation and dissolution in gradient reaction environments: Computational results and stochastic analysis.- Water Resources Research, 44, 10, 1-19. Chaudhuri, A., Rajaram, H. & H. Viswanathan, 2013: Early-stage hypogene karstification in a mountain hydrologic system: A coupled thermohydrochemi-cal model incorporating buoyant convection.- Water Resources Research, 49, 9, 5880-5899. 364 ACTA CARSOLOGICA 44/3- 2015 CONSIDER A CYLINDRICAL CAVE: A PHYSICIST'S VIEW OF CAVE AND KARST SCIENCE Chaudhuri, A., Rajaram, H., Viswanathan, H., Zyvoloski, G. & P. Stauffer, 2009: Buoyant convection resulting from dissolution and permeability growth in vertical limestone fractures.Geophysical Research Letters, 36, 3. Cigna, A.A., 1968: An analytical study of air circulation in caves.- International Journal of Speleology, 3, 1, 41-54. Cigna, A.A., 1993: Environmental management of tourist caves.- Environmental Geology, 21, 3, 173-180. Cooper, M., Perne, M. & M. Covington, 2014: Simple shear stress approximations and a first step in modeling mechanical erosion in caves.- GSA Abstracts with Programs, 46, 534. Covington, M., 2014: Calcite dissolution under turbulent flow conditions: a remaining conundrum.- Acta Carsologica, 43, 195-202. Covington, M., in press: The importance of advection for CO2 dynamics in the karst Critical Zone: an approach from dimensional analysis.- Geological Society of America Special Papers: Caves and Karst Across Time, 516. Covington, M., Gulley, J. & F. Gabrovsek, 2015: Natural variations in calcite dissolution rates in streams: Controls, implications, and open questions.- Geophysical Research Letters, 42, 8, 2836-2843. Covington, M.D., Banwell, A., Gulley, J.D., Saar, M., Willis, I. & C. Wicks, 2012a: Quantifying the effects of glacier conduit geometry and recharge on proglacial hydrograph form.- Journal of Hydrology, 414415, 59-71. Covington, M.D., Luhmann, A.J., Gabrovsek, F., Saar, M.O. & C.M. Wicks, 2011: Mechanisms of heat exchange between water and rock in karst conduits.-Water Resources Research, 47, 10, W10514. Covington, M.D., Luhmann, A.J., Wicks, C.M. & M.O. Saar, 2012b: Process length scales and longitudinal damping in karst conduits.- Journal of Geophysical Research, 117, F1, 1-19. Covington, M.D., Prelovsek, M. & F. Gabrovsek, 2013: Influence of CO2 dynamics on the longitudinal variation of incision rates in soluble bedrock channels: Feedback mechanisms.Geomorphology, 186, 85-95. Covington, M.D., Wicks, C.M. & M.O. Saar, 2009: A dimensionless number describing the effects of recharge and geometry on discharge from simple karstic aquifers.- Water Resources Research, 45, 11, W11410. Culver, D., 2005: Ecotones.-In: D. Culver & W. White (eds.), Encyclopedia of caves, Elsevier, pp. 206-208. Curl, R.L., 1966: Cave conduit enlargement by natural convection.- Cave Notes, 8, 1, 4-8. Dietrich, W.E., Bellugi, D.G., Sklar, L.S., Stock, J.D., Heimsath, A.M. & J.J. Roering, 2003: Geomorphic Transport Laws for Predicting Landscape Form and Dynamics.- Geophysical Monograph, 135, 1-30. Domenico, P. & V. Palciauskas, 1973: Theoretical analysis of forced convective heat transfer in regional ground-water flow.- Bulletin of the Geological Society of America, 84, 12, 3803-3814. Dreiss, S., 1982: Linear kernels for karst aquifers.- Water Resources Research, 18, 4, 865-876. Dreybrodt, W., 1988: Processes in Karst Systems: Physics, Chemistry, and Geology.- Springer, pp. 288, New York, USA. Dreybrodt, W., 1996: Principles of early development of karst conduits under natural and man-made conditions revealed by mathematical analysis of numerical models.- Water Resources Research, 32, 9, 2923. Dreybrodt, W. & D. Buhmann, 1991: A mass transfer model for dissolution and precipitation of calcite from solutions in turbulent motion.- Chemical Geology, 90, 1-2, 107-122. Dreybrodt, W. & F. Gabrovsek, 2000: Dynamics of the evolution of single karst conduits.- In: Speleogen-esis: Evolution of karst aquifers, pp. 184-193. Dreybrodt, W., Gabrov"sek, F. & D. Romanov, 2005a: Processes of Speleogenesis: A Modeling Approach.-ZRC Publishing, pp. 376, Ljubljana, Slovenia. Dreybrodt, W., Gabrovsek, F. & M. Perne, 2005b: Condensation corrosion: a theoretical approach.- Acta Carsologica, 32, 2, 317-348. Dreybrodt, W. & D. Romanov, 2007: Time scales in the evolution of solution porosity in porous coastal carbonate aquifers by mixing corrosion in the saltwater-freshwater transition zone.In: Time in Karst, Postojna, pp. 25-34. Dreybrodt, W., Romanov, D. & F. Gabrovsek, 2002: Karstification below dam sites: a model of increasing leakage from reservoirs.- Environmental Geology, 42, 5, 518-524. Dreybrodt, W., Romanov, D. & G. Kaufmann, 2009: Evolution of isolated caves in porous limestone by mixing of phreatic water and surface water at the water table of unconfined aquifers: A model approach.-Journal of Hydrology, 376, 1-2, 200-208. Dublyansky, V.N. & Y.V. Dublyansky, 2000: The role of condensation in karst hydrogeology and speleogen-esis.- In: A. Klimchouk, D.C. Ford, A. Palmer & W. Dreybrodt (eds.), Speleogenesis: Evolution of karst aquifers. National Speleological Society, pp. 100111, Huntsville, pp. 100-111. ACTA CARSOLOGICA 44/3 - 2015 365 MATTHEW D. COVINGTON & MATIJA PERNE Eisenlohr, L., Kir'aly, L., Bouzelboudjen, M. & Y. Rossi-er, 1997: Numerical simulation as a tool for checking the interpretation of karst spring hydrographs.-Journal of Hydrology, 193, 1-4, 306-315. Emmanuel, S. & Y. Levenson, 2014: Limestone weathering rates accelerated by micron-scale grain detachment.- Geology, 42, 9, 751-754. Fairchild, I.J., Smith, C.L., Baker, A., Fuller, L., Spotl, C., Mattey, D., McDermott, F. & E. I. M. F., 2006: Modification and preservation of environmental signals in speleothems. Earth Science Reviews, 75, 105-153. Farrant, A.R. & P.L. Smart, 2011: Role of sediment in speleogenesis; sedimentation and paragenesis.-Geomorphology, 134, 79-93. Finnegan, N.J., Roe, G., Montgomery, D.R. & B. Hallet, 2005: Controls on the channel width of rivers: Implications for modeling fluvial incision of bedrock.-Geology, 33, 3, 229. Gabrovsek, F., 2007: On denudation rates in karst: O hitrosti denudacije na Krasu.- In: Time in Karst, Postojna, pp. 7-13. Gabrovsek, F. & W. Dreybrodt, 2000: Role of mixing corrosion in calcite-aggressive H2OCO2-CaCO3 solutions in the early evolution of karst aquifers in limestone.- Water Resources Research, 36, 5, 11791188. Gabrovsek, F. & W. Dreybrodt, 2001: A model of the early evolution of karst aquifers in limestone in the dimensions of length and depth.- Journal of Hydrology, 240, 3-4, 206-224. Gabrovsek, F., H'auselmann, P. & P. Audra, 2014: Looping caves versus water table caves: The role of baselevel changes and recharge variations in cave development.- Geomorphology, 204, 683-691. Gabrovsek, F., Menne, B. & W. Dreybrodt, 2000: A model of early evolution of karst conduits affected by subterranean CO2 sources.- Environmental Geology, 39, 6, 531-543. Gabrovsek, F., Romanov, D. & W. Dreybrodt, 2004: Early karstification in a dual-fracture aquifer: The role of exchange flow between prominent fractures and a dense net of fissures.- Journal of Hydrology, 299, 45-66. Geyer, T., Birk, S., Liedl, R. & M. Sauter, 2008: Quantification of temporal distribution of recharge in karst systems from spring hydrographs.- Journal of Hydrology, 348, 3-4, 452-463. Grasso, D. & P.Y. Jeannin, 2002: A Global Experimental System Approach of Karst Springs' Hydrographs and Chemographs.- Ground Water, 40, 6, 608-618. Groves, C.G. & A.D. Howard, 1994: Minimum hydro-chemical conditions allowing limestone cave development.- Water Resources Research, 30, 3, 607615. Gulley, J., Martin, J. & P. Moore, 2014: Vadose CO2 gas drives dissolution at water tables in eo-genetic karst aquifers more than mixing dissolution.- Earth Surface Processes and Landforms, 39, 13, 1833-1846. Gulley, J.D., Martin, J.B., Moore, P.J., Brown, A., Spell-man, P.D. & J. Ezell, 2015: Heterogeneous distributions of CO2 may be more important for dissolution and karstification in coastal eogenetic limestone than mixing dissolution.- Earth Surface Processes and Landforms, 40, 1057-1071. Gulley, J.D., Martin, J.B., Moore, P.J. & J. Murphy, 2013: Formation of phreatic caves in an eogenetic karst aquifer by CO2 enrichment at lower water tables and subsequent flooding by sea level rise.- Earth Surface Processes and Landforms, 38, 1210-1224. Halihan, T. & C.M. Wicks, 1998: Modeling of storm responses in conduit flow aquifers with reservoirs.-Journal of Hydrology, 208, 1-2, 82-91. Hammer, 0., Lauritzen, S.E. & B. Jamtveit, 2011: Stability of dissolution flutes under turbulent flow.- Journal of Cave and Karst Studies, 73, 3, 181-186. Hanna, R. & H. Rajaram, 1998: Influence of aperture variability on dissolutional growth of fissures in karst formations.- Water Resources Research, 34, 11, 2843-2853. Harmon, R.S., Schwarcz, H.P. & D.C. Ford, 1978: Late pleistocene sea level history of bermuda.-Quaternary Research, 9, 2, 205 - 218. Harte, J., 1998: Consider a Spherical Cow.-University Science Books, pp. 283, Sausalito, CA, USA. Heller, E.J. & S. Tomsovic, 1993: Postmodern quantum mechanics.- Physics Today, 46, 38-46. Hoyos, M., Ca~naveras, J.C., S'anchez-Moral, S., Sanz-Rubio, E. & V. Soler, 1998: Microclimatic characterization of a karstic cave: Human impact on mi-croenvironmental parameters of a prehistoric rock art cave (Candamo Cave, northern Spain).- Environmental Geology, 33, March, 231-242. James, E.W., Banner, J.L. & B. Hardt, 2015: A global model for cave ventilation and seasonal bias in spe-leothem paleoclimate records.- Geochemistry, Geophysics, Geosystems, 16, 1044-1051. Jeannin, P.Y. & M. Sauter, 1998: Analysis of karst hydro-dynamic behaviour using global approach: a review.- Bulletin d' Hydrogéologie, 16, 31-48. Kaufmann, G., 2009: Modelling karst geomorphology on different time scales.- Geomorphology, 106, 1-2, 62-77. 364 ACTA CARSOLOGICA 44/3- 2015 CONSIDER A CYLINDRICAL CAVE: A PHYSICIST'S VIEW OF CAVE AND KARST SCIENCE Kaufmann, G. & J. Braun, 2000: Karst aquifer evolution in fractured, porous rocks.- Water Resources Research, 36, 6, 1381-1391. Klimchouk, A., 2007: Hypogene speleogenesis: hydro-geological and morphogenetic perspective. Special Paper 1.- National Cave and Karst Research Institute. Kovacs, A., Perrochet, P., Kiraly, L. & P.Y. Jeannin, 2005: A quantitative method for the characterisation of karst aquifers based on spring hydrograph analysis.- Journal of Hydrology, 303, 1-4, 152-164. Labat, D., Ababou, R. & A. Mangin, 2000: Rainfall-runoff relations for karstic springs. Part I: convolution and spectral analyses.- Journal of Hydrology, 238, 3-4, 123-148. Lamb, M.P., Dietrich, W.E. & L.S. Sklar, 2008: A model for fluvial bedrock incision by impacting suspended and bed load sediment.- Journal of Geophysical Research, 113, F3, 1-18. Lauritzen, S. & J. Lundberg, 2000: Meso-and micro-morphology of caves: solutional and erosional morphology.-In: A. Klimchouk, D. Ford, A. Palmer & W. Dreybrodt (eds.), Speleogenesis: evolution of karst aquifers, National Speleological Society, pp. 408-426. Liedl, R. & M. Sauter, 1998: Modelling of aquifer genesis and heat transport in karst systems.Bulletin d Hydrogéologie, 16, 185-200. Liedl, R., Sauter, M., Huckinghaus, D., Clemens, T. & G. Teutsch, 2003: Simulation of the development of karst aquifers using a coupled continuum pipe flow model.- Water Resources Research, 39, 3, 1-11. Lismonde, B., 2002: Climatologie du Monde Souterrain: Aérologie des systèmes karstiques. Comité Départemental de Spéléologie de l'Isére, pp. 362, Grenoble, France. Long, A.J. & P.C. Gilcrease, 2009: A one-dimensional heat-transport model for conduit flow in karst aquifers.- Journal of Hydrology, 378, 3-4, 230-239. Luetscher, M. & P.Y. Jeannin, 2004: Temperature distribution in karst systems: the role of air and water fluxes.- Terra Nova, 16, 6, 344-350. Luetscher, M., Lismonde, B. & P.Y. Jeannin, 2008: Heat exchanges in the heterothermic zone of a karst system: Monlesi cave, Swiss Jura Mountains.- Journal of Geophysical Research, 113, F2, F02025. Luhmann, A.J., Covington, M.D., Alexander, S.C., Chai, S.Y., Schwartz, B.F., Groten, J.T. & E.C. Alexander, 2012: Comparing conservative and nonconserva-tive tracers in karst and using them to estimate flow path geometry.- Journal of Hydrology, 448-449, 201-211. Luhmann, A.J., Covington, M.D., Myre, J.M., Perne, M., Jones, S.W., Alexander Jr., E.C. & M.O. Saar, 2015: Thermal damping and retardation in karst conduits.- Hydrology and Earth System Sciences, 19, 1, 137-157. Lucic, I., 2011: Interview with Wolfgang Dreybrodt.-Acta Carsologica, 40, 225-232. Maillet, E., 1905: Essais d'hydraulique souterraine et fluviale.- Paris. Mariethoz, G., Kelly, B.F.J. & A. Baker, 2012: Quantifying the value of laminated stalagmites for paleo-climate reconstructions.- Geophysical Research Letters, 39, 5. Milanolo, S. & F. Gabrovsek, 2009: Analysis of Carbon Dioxide Variations in the Atmosphere of Srednja Bijambarska Cave, Bosnia and Herzegovina.-Boundary-Layer Meteorology, 131, 3, 479-493. Montgomery, D.R. & K.B. Gran, 2001: Downstream variations in the width of bedrock channels.- Water Resources Research, 37, 6, 1841-1846. Mylroie, J.E., 2008: Review of the book Hypogene Spe-leogenesis: Hydrogeological and Morphogenetic Perspective.- Journal of Cave and Karst Studies, 70, 129-131. Newson, M., 1971: A model of subterranean limestone erosion in the British Isles based on hydrology.-Transactions of the Institute of British Geographers, 54, 55-70. Opdyke, B.N., Gust, G. & J.R. Ledwell, 1987: Mass transfer from smooth alabaster surfaces in turbulent flows.- Geophysical Research Letters, 14, 11, 11311134. Padilla, A. & A. Pulido-Bosch, 1995: Study of hydro-graphs of karstic aquifers by means of correlation and cross-spectral analysis.- Journal of Hydrology, 168, 73-89. Palmer, A.N., 1991: Origin and morphology of limestone caves.- Bulletin of the Geological Society of America, 103, 1, 1. Palmer, A.N., 2011: Distinction between epigenic and hypogenic maze caves.- Geomorphology, 134, 1-2, 9-22. Perne, M., 2012: Modelling speleogenesis in transition from pressurised to free surface flow.Ph.D. thesis, University of Nova Gorica. Perne, M., Covington, M. & M. Cooper, 2014a: Bedrock channel and cave evolution models based on computational fluid dynamics.- Fall Meeting, AGU, San Francisco, CA, 3-7 Dec. Perne, M., Covington, M.D. & F. Gabrovsek, 2014b: Evolution of karst conduit networks in transition from pressurized flow to free-surface flow.- Hydrology and Earth System Sciences, 18, 4617-4633. ACTA CARSOLOGICA 44/3 - 2015 365 MATTHEW D. COVINGTON & MATIJA PERNE Plummer, L., Wigley, T. & D.L. Parkhurst, 1978: The Kinetics of Calcite Dissolution in CO2 Water Systems at 5° to 60 °C and 0.0 to 1.0 ATM CO2.- American Journal of Science, 278, 179-216. Reimann, T., Geyer, T., Shoemaker, W.B., Liedl, R. & M. Sauter, 2011: Effects of dynamically variable saturation and matrix-conduit coupling of flow in karst aquifers.- Water Resources Research, 47, 1-19. Romanov, D., Kaufmann, G. & W. Dreybrodt, 2008: Modeling stalagmite growth by first principles of chemistry and physics of calcite precipitation.- Geo-chimica et Cosmochimica Acta, 72, 2, 423 - 437. Sanchez-Canete, E.P., Serrano-Ortiz, P., Domingo Pov-eda, F. & A.S. Kowalski, 2013: Cave ventilation is influenced by variations in the CO2-dependent virtual temperature.- International Journal of Speleology, 42, 1-8. Sarbu, S.M. & C. Lascu, 1997: Condensation corrosion in Movila cave, Romania.- Journal of Cave and Karst Studies, 59, 3, 99-102. Serrano-Ortiz, P., Roland, M., Sanchez-Moral, S., Jans-sens, I.A., Domingo, F., Godderis, Y. & A. Kowalski, 2010: Hidden, abiotic CO2 flows and gaseous reservoirs in the terrestrial carbon cycle: review and perspectives.- Agricultural and Forest Meteorology, 150, 321-329. Short, M.B., Baygents, J.C., Beck, J.W., Stone, D.A., Toom-ey, R.S. & R.E. Goldstein, 2005: Stalactite growth as a free-boundary problem: A geometric law and its platonic ideal.- Phys. Rev. Lett., 94, 018501. Siemers, J. & W. Dreybrodt, 1998: Early development of karst aquifers on percolation networks of fractures in limestone.- Water Resources Research, 34, 3, 409-419. Sklar, L.S. & W.E. Dietrich, 2004: A mechanistic model for river incision into bedrock by saltating bed load.- Water Resources Research, 40, 6, 1-22. Spotl, C., Fairchild, I.J. & A.F. Tooth, 2005: Cave air control on dripwater geochemistry, Obir Caves (Austria): Implications for speleothem deposition in dynamically ventilated caves.- Geochimica et Cos-mochimica Acta, 69, 10, 2451-2468. Stafford, K., Land, L. & G. Veni (eds.), 2009: Advances in hypogene karst studies. Symposium 1.- National Cave and Karst Research Institute. Szymczak, P. & A.J. Ladd, 2011: The initial stages of cave formation: beyond the one-dimensional paradigm.-Earth and Planetary Science Letters, 301, 3-4, 424432. Szymczak, P. & A.J.C. Ladd, 2012: Reactive-infiltration instabilities in rocks: fracture dissolution.- Journal of Fluid Mechanics, 702, 239-264. Tarhule-Lips, R.F.A. & D.C. Ford, 1998: Condensation corrosion in caves on Cayman Brac and Isla de Mona.- Journal of Cave and Karst Studies, 60, 2, 84-95. Tobin, B.W., Hutchins, B.T. & B.F. Schwartz, 2013: Spatial and temporal changes in invertebrate assemblage structure from the entrance to deep-cave zone of a temperate marble cave.International Journal of Speleology, 42, September, 203-214. Weisbrod, N., Dragila, M., Nachshon, U. & M. Pillersdorf, 2009: Falling through the cracks: The role of fractures in Earth-atmosphere gas exchange.- Geophysical Research Letters, 36, L02401. Whipple, K.X., Hancock, G.S. & R.S. Anderson, 2000: River incision into bedrock: Mechanics and relative efficacy of plucking, abrasion, and cavitation.-Geological Society of America Bulletin, 112, 3, 490-503. Wigley, T. & M. Brown, 1976: The Physics of Caves.-In: T. Ford & C. Cullingford (eds.), The Science of Speleology. Academic Press, pp. 329-358, New York, pp. 329-358. Wigley, T.M.L. & M. Brown, 1971: Geophysical applications of heat and mass transfer in turbulent pipe flow.- Boundary-Layer Meteorology, 1, 1968, 300320. Wigner, E., 1960: The unreasonable effectiveness of mathematics in the natural sciences.- Communications on Pure and Applied Mathematics, 13, 1-14. Wood, W.W., 1985: Origin of caves and other solution openings in the unsaturated (vadose) zone of carbonate rocks: A model for CO2 generation.- Geology, 13, 822-824. Yanites, B.J. & G.E. Tucker, 2010: Controls and limits on bedrock channel geometry.- Journal of Geophysical Research, 115, F04019, 1-17. 364 ACTA CARSOLOGICA 44/3- 2015