This invention relates to condensation heat transfer.
Typically, condensation happens on solid surfaces that are wetted by the condensing fluid. In this configuration termed filmwise condensation (
This Summary introduces a selection of concepts in simplified form that are described further below in the Detailed Description. This Summary neither identifies key or essential features, nor limits the scope, of the claimed subject matter.
The invention relates to enhancing the condensation heat transfer performance in applications including power generation, thermal management of high-performance electronics, water purification, distillation, natural gas processing, and air conditioning.
Condensation heat transfer can be enhanced via a different mechanism. Instead of utilizing ultra-thin hydrophobic coatings, a hierarchical structure is attached on the condenser surface. This novel hierarchical structure is composed of a thin, highly permeable, thermally conductive porous wick and a highly porous, robust, intrinsically hydrophobic membrane bonded or attached on top of the wick.
In general, a device for providing condensation heat transfer can include a hierarchical structure attached on a condenser surface.
In one aspect, a capillary-driven condensation surface for a condenser surface can include a thermally conductive porous wick and a porous hydrophobic membrane on the wick.
In another aspect, a device having hierarchical structure for attachment to a condenser surface can include a thin, highly permeable, thermally conductive porous wick and a highly porous, robust, intrinsically hydrophobic membrane bonded or attached on top of the wick.
In another aspect, a method of improving a heat transfer coefficient of a thermal system can include placing a capillary-driven condensation surface including a thermally conductive porous wick and a porous hydrophobic membrane on the wick on a surface of a condenser element of the thermal system.
In another aspect, a method of manufacturing a capillary-driven condensation surface for a condenser surface can include placing a porous hydrophobic membrane on a thermally conductive porous wick.
In certain circumstances, the thermally conductive porous wick can be configured to be in thermal contact with the condenser surface.
In certain circumstances, the thermally conductive porous wick can include a sintered metal powder, an electrodeposited porous metal, a metal foam, a metal mesh, a laser-etched metal, a 3D printed metal, a molded surface structure, or a patterned substrate.
In certain circumstances, the thermally conductive porous wick can include a copper foam, a copper mesh, a nickel foam, a stainless steel mesh, or an etched silicon structure.
In certain circumstances, the thermally conductive porous wick can have a porosity of at least 30%.
In certain circumstances, the thermally conductive porous wick can have an average porosity of greater than 30%. In certain circumstances, the thermally conductive porous wick can have an average porosity of less than 98%.
In certain circumstances, the thermally conductive porous wick can have an average pore size of at least 1 micron.
In certain circumstances, the porous hydrophobic membrane can be bonded to or mechanically secured to a surface of the porous wick. For example, the bond or securing can be by physical attachment, such as clamps or ties, thermal attachment, such as by diffusion bonding, or localized melting or solidification, or stress-based attachment, such as by pre-forming the wick and membrane.
In certain circumstances, the porous hydrophobic membrane can have an average pore size of less than 20 microns.
In certain circumstances, the porous hydrophobic membrane can have an average pore size of greater than 10 nanometers.
In certain circumstances, the porous hydrophobic membrane can include an organic polymer or an inorganic material. For example, the organic polymer can be a fluorinated polyolefin or a polyolefin, such as polyethylene, polystyrene, polypropylene (PP), polytetrafluoroethylene, poly(vinylidene fluoride) (PVDF), poly(vinylidene fluoride)-co-hexafluoropropylene (PVDF-HFP), or sulfonated polytetrafluoroethylene. The inorganic material can be a rare earth oxide, silicon oxide, or silicon nitride.
In certain circumstances, the porous hydrophobic membrane can further include a hydrophobic coating. The hydrophobic coating can be a fluorinated alkyl film former, for example, a perfluorinated alkyl silicone alkoxide.
In certain circumstances, the organic polymer can be an electrospun fiber.
In certain circumstances, the electrospun fiber can have an average diameter of between 0.5 microns and 4 microns.
In certain circumstances, the electrospun fiber can have an average diameter of between 0.1 microns and 2 microns.
In certain circumstances, the thermally conductive porous wick can include microchannels.
In certain circumstances, the microchannels can be arranged in rows or bands. The rows or bands can have a spacing between the rows or bands of between 0.25 cm and 10 cm.
In certain circumstances, the rows or bands can be arranged substantially perpendicular to a lengthwise axis of the condenser surface.
In certain circumstances, the porous hydrophobic membrane can include one or more drain ports along a length of the membrane. For example, the drain ports can be adjacent to each of the rows or bands, if present.
In certain circumstances, the method can include securing the capillary-driven condensation surface to the surface of the condenser element.
In certain circumstances, the method can include cleaning a surface of the thermally conductive porous wick prior to placing the porous hydrophobic membrane.
In certain circumstances, the method can include heat treating a surface of the thermally conductive porous wick prior to placing the porous hydrophobic membrane.
The following Detailed Description references the accompanying drawings which form a part this application, and which show, by way of illustration, specific example implementations. Other implementations may be made without departing from the scope of the disclosure.
Enhancing condensation heat transfer can have important implications to reduce local water consumption, reduce CO2 emissions, while simultaneously enhancing the overall cycle efficiency of steam power plants. This is important since most of the electricity produced in the U.S. comes from steam power plants. The state-of-the-art technology, dropwise condensation (DWC), has demonstrated its ability to achieve condensation heat transfer enhancements of up to an order of magnitude. However, it comes short of industrial implementation due to the requirement for thin coatings, which cannot adhere to the necessary industrial timescales required for operation. Therefore, various condensation approaches have been investigated to enhance the durability of condensation enhancing engineered surfaces.
Condensation system, structures and methods described herein have the potential to address concerns of durability by combining thick porous hydrophobic membranes with porous wicking structures on condenser tubes. This concept, termed capillary-driven condensation (CDC), can enhance condensation heat transfer to a level comparable to dropwise condensation according to the preliminary models. Herein, this approach is discussed in depth, and investigate the physics during condensation on the CDC surfaces to understand, design, and develop superior condensation surfaces. Two approaches to developing the surface are described.
The fundamental understanding of the physics behavior of the surface during condensation by fabricating a CDC sample on silicon with highly-defined geometry by modeling, optimization, and rational design of a surface, fabrication of that surface, visualization studies to elucidate the physics, and conduct experimental heat transfer measurements. Vapor condensation of a fabricated structure achieved a heat transfer coefficient ˜240% higher than the theoretical filmwise value at the same operating conditions.
Hierarchical copper surfaces were fabricated with commercially available copper-based foams and meshes, and hydrophobized a copper mesh layer to realize capillary-driven condensation. Capillary-driven condensation was observed on these hierarchical copper surfaces, and measured a 50% heat transfer enhancement over filmwise condensation on a micro-channeled, biphilic, hierarchical copper sample. Modelling predicts a much higher heat transfer enhancement.
Scalable fabrication of a hydrophobic membrane and a porous metal wick was also explored. A parametric study has been conducted to optimize the fabrication recipe of the electrospun fibrous membrane. Sintering and electrodeposition have been studied for fabricating porous copper wick. The heat and mass transfer model predicts >5× heat transfer enhancement over filmwise condensation on an electrospun fiber covered sintered copper powder surface.
Condensation heat transfer can be enhanced via a different mechanism. Instead of utilizing ultra-thin hydrophobic coatings, a hierarchical structure is attached on the condenser surface. This novel hierarchical structure is composed of a thin, highly permeable, thermally conductive porous wick and a highly porous, robust, intrinsically hydrophobic membrane bonded or attached on top of the wick. See, for example,
Condensation heat transfer, whereby a vapor changes phase to liquid by releasing the latent heat, is a critical process in various applications such as steam cycles for power generation, refrigeration cycles for heating, ventilation and air conditioning systems, and heat pipes and vapor chambers for thermal management. Despite decades of research on condenser surface design, state-of-the-art condenser surfaces suffer from low scalability, costly fabrication, and flooding (due to high subcooling). As described herein, condenser surfaces based on the concept of capillary-driven condensation have been designed, fabricated, and tested. By bonding a hydrophobized mesh on top of a highly porous copper foam, the condensate was constrained within the foam layer to form a thin, continuous liquid film. The control of the thickness along with the improved effective thermal conductivity of the liquid film enables us to reduce the thermal resistance: and hence higher heat transfer coefficient. The heat and mass transfer model, which guides surface design, predicts a significant enhancement in heat transfer coefficient compared to the classical filmwise condensation. Experiments are conducted in a custom-built environmental chamber to validate the model prediction. Unlike condenser surfaces designs that rely on sophisticated micro/nanostructuring, the capillary-driven condenser surfaces are easy to fabricate, highly scalable, and can withstand higher subcooling. The insights gained from this work pave a way for enhancing condensation heat transfer in large scale applications.
As shown in
Referring to
The working principle, illustrated in
The benefits of a thin, high-permeability, high-thermal conductivity, porous structured wick to enhancing condensation heat transfer are manifold. The wick-condensate composite has a higher effective thermal conductivity which reduces the thermal resistance of the condensate film. Minimizing the height of this layer further reduces its thermal resistance as the heat transport distance decreases. Similarly, the benefits of an intrinsically hydrophobic membrane are manifold. Unlike those highly degradable, self-assembled monolayer hydrophobic coatings, intrinsically hydrophobic membranes are robust and presently used in the water treatment/desalination industries. Most importantly, the micrometer-sized pores of the hydrophobic membrane could generate a significant capillary pressure, which acts as the driving force for the condensate to overcome viscous pressure loss along the wicking structures and flow towards the designated exit ports. Several parameters can be carefully designed in order to optimize the condenser's heat transfer performance, such as: membrane pore size, membrane thickness, porosity of the membrane, wick thickness, permeability of the wick, porosity of the wick.
The present approach can achieve much larger condensation heat transfer performance than conventional extended surfaces, which are commercially available.
The present approach can overcome the issue of robustness and durability of ultra-thin hydrophobic coatings in dropwise condensation by utilizing components, which are individually robust. The wicking structure can be made from a robust metal foam, and the porous hydrophobic membrane can be a commercially available polymer membrane thick enough to be robust. The performance can be comparable or higher than dropwise condensation depending on the surface design.
The driving force for condensate flow in the approach, i.e., the capillary pressure generated by the membrane pores, is completely passive and can be enormous (>10 MPa) by utilizing a small membrane pore (on the order of 1 μm), which makes the approach superior than those previously reported wicking condensation surfaces that either use external pumping force (see reference 8) or gravitational force (see reference 9) to drive the condensate flow:
Some recent studies have applied hydrophobic coatings on the top portion of a wicking structure in order to retain the condensed liquid inside the wicking structure with the capillary pressure generated at the hydrophobic top. These partially-hydrophobically coated (i.e., amphiphilic) wicking structures include nanowires (see reference 10), inverse opals (see reference 10), anodic aluminum oxide (AAO) membranes (see reference 9), micropillar arrays (see reference 9). However, these amphiphilic wicking structures couple the capillary driving pressure generated by the hydrophobic top layer and the viscous pressure loss generated by the hydrophilic bottom layer since the two layers share the same pore geometry, limiting the design space and potential for enhancing heat transfer performance. In addition, many of these wicking structures (e.g., nanowires and micropillar arrays) require expensive, lab-scale fabrication, limiting scale-up applications. Moreover, the ultra-thin hydrophobic coatings utilized by the above amphiphilic structures are non-robust and would not endure industrial-level operations even for weeks. The approach fundamentally surpasses the existing approaches mentioned above by decoupling the capillary driving force and the viscous pressure loss with a two-layer hierarchical surface design. By adding an intrinsically hydrophobic porous membrane on top of the wicking structure, the pore sizes of the wicking structures can be independently tailored from those of the membrane layer to achieve a lower viscous pressure drop inside the wicking structures to form a highly permeable wick while maximizing the capillary driving pressure by reducing the size of the membrane pores. These physics can be modeled, and rational designs can be made. Furthermore, low cost, robust, and scalable fabrication techniques can be used, such as electrospinning fiber membranes and electroless deposition of porous metals.
The proposed method, systems and surfaces have applicability in steam power generation. Steam power plants account for ˜60% (see reference 10) of the total electricity generated in the US and ˜80% in the world. Retrofitting existing power plants, is a huge market for technologies that promise enhancements in condensation heat transfer that can last on the order of 15 years or more.
Moreover, the thermal management of high-performance electronics sector utilizes heat pipes and vapor chamber technologies, which can be found in conventional laptops, cell phones, and desktops. Heat pipes and vapor chambers both contain evaporators and condensers. A proof-of-concept condenser utilizing this approach is described herein. It is fabricated on silicon utilizing micromachining technology which is common to the industry (see
As described herein, a device for providing condensation heat transfer can include a hierarchical structure attached on a condenser surface.
Referring to
Capillary-driven condensation, condensation occurs on the hydrophilic micro/nanostructured wick, and the condensate is then forced out due to the capillary pressure buildup at the menisci formed in the porous hydrophobic membrane. The presence of the structures and the resulting capillarity helps maintain a stable liquid film while driving liquid flow. By tailoring the size of the pores in the membrane and the geometry of the wicking structure, the capillary pressure generated can be maximized and the flow rate of the condensate can be optimized to increase the rate of condensation that the wicking structure can support.
In one aspect, a capillary-driven condensation surface for a condenser surface can include a thermally conductive porous wick and a porous hydrophobic membrane on the wick.
In another aspect, a device having hierarchical structure for attachment to a condenser surface can include a thin, highly permeable, thermally conductive porous wick and a highly porous, robust, intrinsically hydrophobic membrane bonded or attached on top of the wick.
For example, a robust hydrophobic membrane can be combined with a porous, high-thermal-conductivity metal wick and wrapped around the external surface of the condenser tube. During the condensation process, water vapor transports through the membrane pores and condense inside the porous metal wick, forming a thin film with a thickness constrained by the capillary pressure generated at the base of the membrane pore. Condensed water drains out through designated exit port to avoid a flooding issue. The drain can be by gravity or by pump or other pressure differential.
In another aspect, a method of improving a heat transfer coefficient of a thermal system can include placing a capillary-driven condensation surface including a thermally conductive porous wick and a porous hydrophobic membrane on the wick on a surface of a condenser element of the thermal system.
In another aspect, a method of manufacturing a capillary-driven condensation surface for a condenser surface can include placing a porous hydrophobic membrane on a thermally conductive porous wick.
Surprisingly, vapor condensation of a fabricated structure can achieve a heat transfer coefficient 50% to 500% higher than the theoretical filmwise value at the same operating conditions. Moreover, capillary-driven condensation was observed on these hierarchical copper surfaces, and measured a 50% heat transfer enhancement over filmwise condensation on a micro-channeled, biphilic, hierarchical copper sample.
In certain circumstances, the thermally conductive porous wick can be configured to be in thermal contact the condenser surface.
In certain circumstances, the wick can have thickness of between 5 microns and 100 microns, for example, 5 microns, 10 microns, 15 microns, 20 microns, 25 microns, 30 microns, 35 microns, 40 microns, 45 microns, 50 microns, 55 microns, 60 microns, 65 microns, 70 microns, 75 microns, 80 microns, 85 microns, 90 microns, 95 microns, or 100 microns.
In certain circumstances, the thermally conductive porous wick can include metal wire cloth, perforated sheets, and metal foams, for example, a sintered metal powder, an electrodeposited porous metal, a metal foam, a metal mesh, or a patterned substrate.
In certain circumstances, the thermally conductive porous wick can include a copper foam, a copper mesh, a nickel foam, a stainless steel mesh, or an etched silicon structure, for example, silicon nitride. The wick can include partially-hydrophobically coated (i.e., amphiphilic) wicking structures include nanowires, inverse opals, anodic aluminum oxide (AAO) membranes, or micropillar arrays. The wick can include perforated stainless steel, perforated brass, perforated steel, perforated aluminum, perforated copper, stainless steel mesh, brass mesh, steel mesh, aluminum mesh, copper mesh, copper foam, or nickel foam, sintered spherical metal powders with diameter <10 μm, spherical metal powders with diameter <50 μm, or dendritic metal powders with size <45 μm, or combinations thereof.
As described in more detail below, copper-based foams and meshes, silicon and silicon nitride structures, and a hydrophobized copper mesh layer have been shown experimentally to achieve capillary-driven condensation. Sintering and electrodeposition have been studied for fabricating porous copper wick. The heat and mass transfer model predicts >5× heat transfer enhancement over filmwise condensation on an electrospun fiber covered sintered copper powder surface.
In certain circumstances, the thermally conductive porous wick can have a porosity of at least 30%, at least 35%, at least 40%, at least 45%, at least 50%, at least 55%, at least 60%, at least 65%, at least 70%, at least 75%, at least 80%, at least 85%, at least 90%, or at least 95%. The thermally conductive porous wick can have a porosity of no greater than 98%.
In certain circumstances, the thermally conductive porous wick can have a thickness of no more than 500 microns, for example, no more than 1 micron, 5 microns, 10 microns, 20 microns, 30 microns, 50 microns, 70 microns, 90 microns, 100 microns, 150 microns, 200 microns, 250 microns, 300 microns, 350 microns, 400 microns, 450 microns, or 500 microns.
In certain circumstances, the thermally conductive porous wick can have an effective thermal conductivity between 0.6 W/mK and 400 W/mK. The effective thermal conductivity can be selected based on the thickness of the wick. For example, if the wick has thickness below 20 microns, then an effective thermal conductivity of above 0.6 W/mK should give some enhancement and if the wick has a thickness above 100 microns, then an effective thermal conductivity of above 10 W/mK can be needed to get some heat transfer enhancement. The higher the effective thermal conductivity the better.
In certain circumstances, the thermally conductive porous wick can have an average pore size of less than 30 microns, less than 25 microns, less than 20 microns, less than 18 microns, less than 16 microns, less than 15 microns, less than 14 microns, less than 13 microns, less than 12 microns, less than 11 microns, or less than 10 microns.
In certain circumstances, the thermally conductive porous wick can have an average pore size of at least 1 micron, at least 2 microns, at least 3 microns, at least 4 microns, at least 5 microns, at least 6 microns, at least 7 microns, at least 8 microns, at least 9 microns, or at least 10 microns.
In certain circumstances, the porous hydrophobic membrane can be bonded to or mechanically secured to a surface of the porous wick. For example, the bond or securing can be by physical attachment, such as clamps or ties, thermal attachment, such as by diffusion bonding, or localized melting or solidification, or stress-based attachment, such as by pre-forming the wick and membrane.
In certain circumstances, the porous hydrophobic membrane can have an average pore size of less than 20 microns, less than 18 microns, less than 16 microns, less than 15 microns, less than 14 microns, less than 13 microns, less than 10 microns, less than 8 microns, less than 6 microns, less than 5 microns, less than 2 microns, less than 1 micron, less than 0.5 microns, less than 0.25 microns, or less than 0.1 microns.
In certain circumstances, the porous hydrophobic membrane can have an average pore size of greater than 10 nanometers, greater than 50 nanometers, greater than 100 nanometers, greater than 200 nanometers, or greater than 500 nanometers.
In certain circumstances, the porous hydrophobic membrane can have a thickness of about 0.1 microns, 0.25 microns, 0.5 microns, 1.0 microns, 1.5 microns, 2.0 microns, 2.5 microns, 3.0 microns, 3.5 microns, 4.0 microns, 4.5 microns, 5.0 microns, 6.0 microns, 7.0 microns, 8.0 microns, 9.0 microns, or 10 microns. In certain embodiments, the porous hydrophobic membrane can have a thickness of 10 mm or less.
In certain circumstances, the porous hydrophobic membrane can include an organic polymer or an inorganic material. For example, the organic polymer can be a fluorinated polyolefin or a polyolefin, such as polyethylene, polystyrene, polypropylene (PP), polytetrafluoroethylene, poly(vinylidene fluoride) (PVDF), poly(vinylidene fluoride)-co-hexafluoropropylene (PVDF-HFP), or sulfonated polytetrafluoroethylene. The inorganic material can be a rare earth oxide, silicon oxide, or silicon nitride.
In certain circumstances, the porous hydrophobic membrane can further include a hydrophobic coating. The hydrophobic coating can be a fluorinated alkyl film former, for example, a perfluorinated alkyl silicon alkoxide.
In certain circumstances, the organic polymer can be an electrospun fiber. The fiber can be a nanofiber or a microfiber. For example, an electrospun fibrous membrane can be deposited on a wick at the time of formation. In certain circumstances, the wick can receive a solvent cleaning treatment or plasma cleaning treatment, or both, to enhance the attachment between the electrospun fibers and to the wick.
In certain circumstances, the electrospun fiber can have an average diameter of between 0.5 microns and 4 microns. For example, the electrospun fiber can have an average diameter of about 0.1 microns, 0.2 microns, 0.4 microns, 0.6 microns, 0.8 microns, 1.0 microns, 1.2 microns, 1.3 microns, 1.4 microns, 1.5 microns, 1.6 microns, 1.7 microns, 1.8 microns, 1.9 microns, or 2.0 microns.
In certain circumstances, the thermally conductive porous wick can include microchannels. The microchannels can have a width of about 0.5 microns, 0.6 microns, 0.7 microns, 0.8 microns, 0.9 microns, 1.0 microns, 1.2 microns, 1.3 microns, 1.4 microns, 1.5 microns, 1.6 microns, 1.7 microns, 1.8 microns, 1.9 microns, or 2.0 microns. The microchannels can be spaced by about 1.00 micron, 1.25 microns, 1.50 microns, 1.75 microns, 2.00 microns, 2.25 microns, 2.50 microns, 2.75 microns, 3.00 microns, 3.5 microns, 4.0 microns, 4.5 microns, 5.0 microns, 5.5 microns, 6.0 microns, 6.5 microns, 7.0 microns, 7.5 microns, 8.0 microns, 8.5 microns, 9.0 microns, 9.5 microns, or 10.0 microns. In other circumstances, spacing between microchannels can be about 0.2 cm, 0.4 cm, 0.6 cm, 0.8 cm, 1.0 cm, 1.2 cm, 1.4 cm, 1.6 cm, 1.8 cm, 2.0 cm, 2.2 cm, or 2.4 cm.
In certain circumstances, wicking lengths can be about half the circumference of the condenser tubes.
In certain circumstances, the microchannels can be arranged in rows or bands. The rows or bands can have a spacing between the rows or bands of between 0.25 cm and 10 cm. A drainage gap can be formed between each row or band. The row or band can be secured to the condenser surface by shrinkage, elastic behavior, by stitching, by physical clamp, such as using a shrink wrap screen. For example, the rows or bands can have a spacing between the rows or bands of about 0.25 cm, 0.50 cm, 0.75 cm, 1.00 cm, 1.25 cm, 1.50 cm, 1.75 cm, 2.00 cm, 2.25 cm, 2.50 cm, 2.75 cm, 3.00 cm, 3.5 cm, 4.0 cm, 4.5 cm, 5.0 cm, 5.5 cm, 6.0 cm, 6.5 cm, 7.0 cm, 7.5 cm, 8.0 cm, 8.5 cm, 9.0 cm, 9.5 cm, or 10.0 cm.
In certain circumstances, the rows or bands can be arranged substantially perpendicular to a lengthwise axis of the condenser surface. In certain circumstances, the rows or bands can be at an angle relative to the perpendicular, for example, 5, 10, 15, 20, 25, 30, 35, 40, or 45 degrees from perpendicular.
In certain circumstances, the porous hydrophobic membrane can include one or more drain ports along a length of the membrane. For example, the drain ports can be adjacent to each of the rows or bands, if present. The drain ports can be spaced apart by about 0).25 cm, 0.50 cm, 0.75 cm, 1.00 cm, 1.25 cm, 1.50 cm, 1.75 cm, 2.00 cm, 2.25 cm, 2.50 cm, 2.75 cm, 3.00 cm, 3.5 cm, 4.0 cm, 4.5 cm, 5.0 cm, 5.5 cm, 6.0 cm, 6.5 cm, 7.0 cm, 7.5 cm, 8.0 cm, 8.5 cm, 9.0 cm, 9.5 cm, or 10.0 cm in a row along a bottom edge of the surface.
In certain circumstances, the method can include securing the capillary-driven condensation surface to the surface of the condenser element. Securing can be achieved by physical attachment, such as clamps or ties, thermal attachment, such as by diffusion bonding, using adhesive, or localized melting or solidification, or stress-based attachment, such as by pre-forming the wick and membrane on the substrate.
Without being bound to any particular theory, there can be vapor transport in membrane pores during operation of a CDC surface. Vapor transport in membrane pores can be modelled. As described herein, CDC operates better when the vapor transport resistance is minimized or eliminated, such as with thinner membranes or with liquid (rather than vapor) transport in the pores. Given that fluid is expected to condense inside of the wicking structure during the initial nucleation stage, vapor transport is expected to occur in the membrane pores. Furthermore, vapor transport is expected when the meniscus of condensate is pinned at the bottom of the membrane pore at the wick-membrane interface. Therefore, it is important to understand and model the physics of vapor transport in the membrane. Nevertheless, the type of transport in the pore can be a function of the wettability of the membrane and its geometry. Here, cylindrical geometry for the membrane pores can be assumed as it simplifies the model. Moreover, the geometry can be made with the fabrication method. Therefore, it is possible to both model and fabricate with the same geometry. The model is developed below along with discussion on the effect of saturated vapor operating conditions on the modelling approach.
To begin, the operating conditions of a power plant condenser, which are established by the saturation temperature and pressure of the steam, determine the type of flow that will occur through membrane pores. The vapor can flow in two major regimes. First, the vapor molecules can travel through a pore with minimal collisions between molecules, and most collisions with the membrane pore walls. This type of flow is termed free-molecule flow or Knudsen flow. In the other regime, molecule-molecule collisions dominate over wall-molecule collisions. This is called viscous flow, and the equations for this type of flow are well-developed. There is a regime called the transition regime which considers molecule behaviour from both the Knudsen and the Poiseuille flow regime. To determine which regime is applicable, the mean free path is compared against the pore diameter. The mean free path is the average distance that a molecule travels before it collides with another molecule. The mean-free path is given by the following relation.
where kB is the Boltzmann constant (1.3806485×10−23 J/K), T is the absolute temperature, P the average pressure within the membrane pores, and dc is the collision diameter for water molecules (2.641×10−10 m). The ratio between the mean free path and the pore diameter is called the Knudsen number, and it gives a measure of the flow regime, Kn=λ/dp. If Kn>10 there is free-molecule or Knudsen flow. If Kn<0.01, Poiseuille flow ensues. If conditions are in between, 0.01<Kn<10, a transition regime is applicable where descriptions from both types of flow should be included.
Many models have been developed to describe fluid flow through porous media. Among them is the well-known dusty-gas model. This model can capture free-molecule flow, viscous flow, and continuum diffusion as part of its development. The system consists of a single, pure vapor of water molecules. Equations for this type of flow simplify significantly for this case.
In this analysis, a steam saturation temperature of 35° C. was assumed which is not uncommon near the condenser in steam power plants. The resultant mean free path is around 2.4 μm. Thus, different flow types can occur in membrane pores if the pore size is much larger, or much smaller, than 2.4 μm. In this analysis, pore diameters up to 30 μm were considered. For this operating condition, pore diameters below 250 nm can be said to be in the Knudsen flow regime and those above in the transition regime. For the transition regime, both the Knudsen and the Poiseuille flow contribution were considered.
When condensation occurs, a low-pressure region is created and flow is directed toward the condensing surface. The pure vapor pressure-driven flux in the transition region can be described by
where R is the gas constant (8.3144598 J/mol/K), T is the absolute temperature, Δp is the pressure drop across the membrane, tm is the thickness of the membrane, and
where DiK is the Knudsen diffusion coefficient,
where Ko is the Knudsen flow parameter or permeability coefficient in free molecule or Knudsen flow regime, and
It is worth noting the transport coefficients and proportionality constants depend on the structure of the porous medium. For the simple case of a porous membrane consisting of cylindrical pores of length tm and diameter dp the Knudsen and Poiseuille flow parameters are simply related as Ko=dp/4 and Bo=dp2/32 respectively. Thus, the mean flow permeability becomes,
where tm is the membrane thickness, R is the gas constant, T is the absolute temperature,
The flux through a pore can now be calculated by multiplying the mean permeability through a pore by the pressure across the membrane
where Pvap is the vapor pressure, and Psat (Twall+ΔTwick) is the liquid saturation pressure at the wick-membrane interface, here used as an approximation to the saturation pressure right outside the liquid-vapor interface at the bottom of the pore. The units of equation 6 are moles/m2/s. To convert this to a mass flow rate, one can multiply by the molar mass M in kg/mol. This quantity refers to the mass flux in the area of one pore, so one can multiply by the total frontal surface area AT and the membrane porosity ϕm to obtain the total mass flow rate through the membrane. The mass flow rate through the pore is then given by,
Assuming a square shape to the frontal condensing surface of length L, AT=L2, and if L/4twick>>1, or if the wick sides are covered by an impermeable material, then condensation on the sides of the wick can be neglected. Maximizing the membrane mass flux is crucial. Since the membrane determines both the maximum possible driving capillary pressure available, and the mass flow capacity through the frontal area, which are both limited by the pore diameter, optimizing the pore diameter and maximizing the porosity is crucial.
An enthalpy balance yields the following equation, accounting for condensation in each pore through the frontal surface area,
Heat flux through the wick—assuming this is constant through the frontal area of the wick, which is not necessarily true since the membrane is on the frontal surface and the open pore area transfers heat by phase-change but the solid area is assumed to insulate—is given by
Here, keff is the effective thermal conductivity of the wick, and ΔTwick is the temperature drop across the wick. Finally, one can describe flow through the structured wick by first utilizing Darcy's law,
where κ is the structured wick permeability, μl is the liquid dynamic viscosity, ρt is the liquid density, g the gravitational constant, and dp/dx the additional driving pressure gradient from capillarity.
Adding a membrane atop the wicking structure can promote additional fluid driving pressure by capillarity as well as a vapor transport resistance through the thickness of the membrane. With careful design of the membrane, one can find a trade-off where it may be optimized to maximize the additional capillary driving pressure, while minimizing resistance to flow. The additional capillary-driving pressure produced can be approximated by the following expression for cylindrical pores, which represents the maximum capillary pressure that can be sustained by a pinned meniscus, or the burst pressure to enter a pore,
where Pcap,max is the pore-entry burst pressure, P0 is the pressure of the vapor above the interface, PL is the pressure of the liquid just below the interface, σ is the liquid surface tension, CA is the contact angle of the hydrophobic membrane, and dp is the radius of curvature of the liquid-vapor interface emanating from the pore, assumed to be the pore diameter at the burst pressure. A review of key variables is shown below in Table 1.
Equations 1 through 13 form a complete picture of the preliminary model which considers vapor flow in the transition regime. This model describes the expected behaviour from CDC surfaces when condensation happens with vapor transport in the pores. Despite of the type of transport in the pores, whether liquid or vapor, the transport in the wicking structure of the model continues to apply. To describe liquid transport in the pores, the fluid mechanics in the pore can be coupled with the Schrage equation (or the moment-method) for condensation. The overall trends remain similar. Reasonable values for multiple variables can be prescribed in the model and sweep others to understand trade-offs in design and the expected heat transfer during condensation. The goal was to obtain the maximum possible heat transfer performance while avoiding surface flooding, which degrades heat transfer. Based on modelling, it was found that wicking structures can have high permeability as well as effective thermal conductivity. However, it was found that thermal conductivity becomes less important for higher permeabilities. Thus, designing for higher permeabilities becomes a priority. The model highlights bounding values for the design. Moreover, to optimize heat transfer performance, pore diameters can be below 20 μm. Smaller optimized pore diameters yield higher capillary pressure and heat transfer performance. Note that the hydrophobic membrane was here assumed to be a cylindrical through-pore membrane, but this geometry is not necessarily required for operation.
The preliminary model showed us that it was advantageous to design for a thin wicking structure with high thermal conductivity and high porosity, utilize the thinnest membrane possible with the smallest pores that could sustain the liquid thin film below designed level and avoid flooding at a given operating conditions. For brevity, those results are not presented here, but rather a finite element heat transfer coefficient model is presented below: This model is built utilizing geometry which is directly related to the structure fabricated in the examples described below, one characterized by highly-defined geometry.
The previous section introduced an analytical model which was helpful to get initial estimates of the characteristic length scales required for the hierarchical structured CDC surface and their expected performance. Moreover, that model revealed general trends on how such surfaces operate. The geometry of either the wick or the membrane determine the magnitude of their associated heat and mass transfer resistances, which can be independently designed and optimized to enhance heat transfer. In this section, a model is utilized to rationally design the CDC surfaces utilizing highly defined geometry. This will allows for the tailoring of designs to attain the highest HTC possible at a given operating condition, while simultaneously avoiding top-surface flooding. An explanation of how flooding occurs on these surfaces in the visualization studies is described below. COMSOL is utilized to calculate both the permeability and the heat transfer coefficient (HTC) for micro-pillared wicks as a function of the geometry of the hydrophobic membrane and the wick. Later versions of this model utilize analytical permeability results for micropillar wicks from Byon and Kim (cited above). The results from the HTC model are combined with analytical porous media models, and mechanical stability calculations, to find the flooding criteria, and mechanically stable surfaces, respectively. From the results, designs can be developed.
For the HTC, a conduction model was utilized neglecting convection effects due to a small Jakob and Peclet numbers in the condensate flow direction of the wick. The HTC is calculated on a micro-pillar unit cell (
For vapor transport through the membrane pores, the dusty-gas model from the previous section can be used and the previous equation updated. Previously, the equation was limited to large ratios of the thickness of the membrane tm to the pore radius rp (or very thick membranes). In this case, the transmission probability—the probability that a molecule entering the pore from the top surface will make it all the way across the membrane-simplified to η≈2.67(rp/tm). However, as the membrane thickness decreases and the ratio is close to unity or a little less, this assumption is no longer valid. Thus, the transmission probability was modified after the equations provided by Berman, A. S. Free molecule transmission probabilities. J. Appl. Phys. (1965), which is incorporated by reference in its entirety. First, a reduced length, L=tm/rp, is introduced for a cylindrical pore. Then, the transmission probability across the pore is approximated by equation 14 below. This equation is then utilized in the dusty-gas model for vapor transport in the transition regime.
Next, the HTC was calculated, viscous pressure-drop across the structure for a given subcool was calculated, and the stressed induced on the membrane to predict possible membrane failure was estimated.
where σl is the distributed load, L is the longest length connecting two pillars in a unit cell, E is the Young's Modulus, and I is the moment of inertia. The moment of inertia for the assumed geometry is given by equation 16,
where d is the pillar diameter representing the beam width, and tm is the membrane thickness. The maximum bending stress at the center is then given by equation 17.
The distributed load is overestimated and assumed to be the highest local Laplace pressure generated in a superhydrophobic pore which is 4σ/dp. In reality, the pore is not superhydrophobic and this value is scaled down due to a lower contact angle. The smallest pore in this study (1 μm) yields a local pressure of 0.28 MPa for the overestimated value. From this load one can calculate the bending stress for different unit cell geometries and compare it to the available measurements of the tensile strength for the materials and geometries utilized. A quick calculation utilizing a wick geometry with d=7 μm, d/l=0.5, and membrane made of Si3N4 with geometry with dp=1 μm and tm=8 μm, reveals that the maximum bending stress is 3.79 GPa, which is close to the 14.1 GPa tensile strength estimated from Yoshioka, T., et al., Sensors Actuators. A Phys. 82, 291-296 (2000), which is incorporated by reference in its entirety. Smaller thicknesses for the membrane utilizing this model yield higher bending stresses. However, values for the principal stresses which are representative of the geometry here modelled are required to obtain better estimates. Finite element simulations of the first principle stress were computed using the fabricated geometry with COMSOL Multiphysics. The results show that that the first principle stress is much lower (˜10×) than the expected tensile strength for a 500 nm membrane. Therefore, the membrane was expected to uphold the required capillary pressures from the constrained fluid layer in the wick. Despite this result, each designed geometry should be checked for failure, especially as the membrane thickness decreases and the pillar pitch increases. Nevertheless, the main failure mechanism can be pinpointed. Notably, the largest bending moment varies as ˜L4 which is a strong dependence on the pillar spacing. Thus, to reduce the bending stress and achieve thinner membranes, one needs to consider smaller pillar to pillar spacing.
The modelling in this section helps us understand the relationship between the geometry in the wicking structure and that of the membrane. Specifically, the model shows how the geometries affect the HTC, flooding criteria, and mechanical stability of the membrane. The next section details the fabrication for the structures which were rationally selected utilizing a model with the framework shown in this section.
An approach to fabrication for the capillary-driven condensation structure is shown in
Initial fabrications were done at a millimeter scale sample in order to test the feasibility of the proposed fabrication approach. It was found that it was possible to generally achieve the desired structure and thus showed the feasibility of the approach. The next phase was scaling up the design to the centimetre scale to demonstrate the fabrication at industrially relevant length scales. In this attempt, it was difficult to achieve a large-scale printing of the pattern onto the photoresist utilizing direct-write lithography due to the large size of the drawing. The design was ordered in a photomask from a company for used with a UV light exposure tool. However, this approach did not yield good results due to spatial variations in the exposure dose at these critical dimensions (˜1 μm feature size). In this section, the MLA-150 direct write lithography tool was used and a feature workaround was utilized to print a large drawing successfully. Previously, it was not possible to print large patterns due to the high density of the drawing caused by the many circles that needed to be converted into the format of the tool. However, a feature on the tool was found that could be used as a workaround. Briefly, a smaller millimetre-sized unit cell of the structure was copied in space and “stitched” together to achieve centimetre scale structures. Smaller drawings could successfully be converted to the tool's format, whereas larger drawings could not. Thus, a pattern “stitching” approach was utilized to fabricate the centimetre-scale sample. Fabricating one design at a time then became a feasible approach rather than printing multiple designs on large areas of a 6-inch wafer using either direct-write lithography or a photomask with UV exposure.
In order to achieve capillary-driven condensation with the fabricated hydrophilic structure, simple and effective methods to coat the membrane as hydrophobic for condensation tests were investigated. For the greatest simplicity, simply coating the entire sample hydrophobic in a one step method can be used. After all, nucleation is expected to occur everywhere on a subcooled surface. Therefore, it was believed that nucleation would occur inside of the wicking structure, and that nucleates would grow in the structure and connect as a continuous condensate film. However, this was not the case. Although it is the easiest coating method, this configuration turns out to be neither advantageous nor desired. Thus, instead of coating the entire structure hydrophobic, a method was developed to coat only the top surface of the membrane as hydrophobic. This way, a bi-philic structure would be formed where part of the membrane was hydrophobic while the rest of the structure remained hydrophilic. In this configuration, the wicking layer can fill with condensate.
To selectively coat the structure, areas that should not be coated hydrophobic were protected.
A horizontal test rig was built on an optical microscope and perform visualization studies of condensation in atmospheric-air conditions at low condensation intensities. The purpose was to study the interaction of condensate with the biphilic surfaces.
Many visualization studies were carried out with various coating techniques. Below, results utilizing samples coated with the FAS coating recipe are shown. To begin the visualization study, the structures were coated in two ways. First, and for ease, the first large-area samples were coated to be entirely hydrophobic. This means that both the membrane as well as the wick were fully hydrophobic (
First, it was observed that some nucleation appears on the top surface of the sample (left image). As more condensation occurs, it was observed that nucleation and coalescence of liquid droplets occurs inside the wicking structure (center image). Here, top-surface droplets were absorbed through the membrane pores via liquid bridging between the meniscus pinned at the pore and nearby droplets that can make contact and bridge with the meniscus. Moreover, some droplets remain on the top surface of the membrane and grow. The right-side image shows the formation of small condensate islands inside the wick rather than a continuous film. Further, the incipience of flooding was observed. This occurs in the form of ejected or ‘bursting’ of droplets from the top surface of the membrane. To explain this phenomenon, it was noted that the rough hydrophobic wick increases the contact angle of liquid inside the wick, which increases the viscous pressure drop and causes condensate islands to release their pressure by flooding or “bursting” liquid through nearby membrane pores. One down side of these “bursting” droplets is that they strongly pin on the membrane surface. From these observations, one can expect that the heat transfer would be poor since the pinned droplets increase the overall thermal resistance during condensation. Moreover, below the membrane there are large air-filled sections without condensate islands which lower the effective thermal conductivity of the wick. In addition, these regions show no condensate nucleation meaning that heat transfer is not by phase change. Thus, this form of condensation is not advantageous for heat transfer.
It is worthy to note that naturally filling the wick with a hydrophobic sample may be possible with taller wicking structures that are less prone to “bursting” droplets. However, such a configuration may lower the HTC. Nevertheless, it was concluded that coating the entire structure hydrophobic is not a preferred configuration in capillary-driven condensation with the present sample geometry.
Having seen these results, methods to selectively coat the structure were sought, such that the micropillar structure at the bottom of the membrane would remain hydrophilic after fabrication and fill naturally during condensation. It was found that the selectively coated surface not only gets rid of top surface droplets on the membrane by continually absorbing them into the wicking structure through the membrane pores, but it also avoids flooding (
Visualization experiments comparing the fully hydrophobic samples with the selectively coated samples reveal superior performance and understanding of the possible operating modes of CDC.
In this section, successful heat transfer measurements of condensation in both a bare copper surface and the engineered CDC sample attached to the copper block are reported. A saturated vapor environment at ˜60° C. was created. To measure the heat flux, five thermocouple readings were utilized along the copper rod at known distances. When the profile is linear at steady state, a data point was taken for ˜10 mins or longer.
In conclusion, the silicon-based CDC surfaces with highly defined geometry during condensation in an air-ambient were modelled, designed, fabricated, functionalized, and tested. Most importantly, measurements of the heat transfer in pure-vapor were taken. It was found that the surfaces can achieve enhancements in the HTC of up to ˜240% (in a 60° C. saturated vapor environment) compared to the theoretical filmwise condensation prediction for the sample with a sparse micropillar spacing of 150 μm. Further, higher HTCs were expected for the samples with closer micropillar spacing. Heat transfer can also be studied with smaller micropillar pitch values. Overall, the samples were seen to lower the temperature drop at the surface or the subcool, which indicated they are lowering the thermal resistance. These results highlight the promise of CDC surfaces in achieving high HTCs. In order to show their promise in a more scalable fashion, the next section concerns the development of scalable CDC surfaces utilizing copper and other scalable materials.
For achieving capillary-driven condensation on a large scale, i.e., for application in power plant condensers, the vision is to manufacture a hierarchical condenser tube as shown in
The hydrophobic membrane is a core component of the capillary-driven condenser design, which provides Laplace pressure from the outward-curving liquid-vapor interfaces to drive the condensate transport into the wick structures below the membrane. There are many hydrophobic membrane materials that are commercially available, such as polytetrafluoroethylene (PTFE), polypropylene (PP) and polyvinylidene difluoride (PVDF). A list of some commercially available choices for hydrophobic membranes and their corresponding properties are shown in Table 2.
PTFE membranes are intrinsically hydrophobic, chemically stable, and applicable to large scale industrial applications. They have a relatively wide range of pore size (0.1-10 μm) and thickness that are commercially available, and their prices vary from several hundred to several thousands of dollars per meter squared depending on the geometry. The unlaminated PTFE membranes do not have a supportive layer, while the laminated ones have a supportive layer of PP or polyester (PE) to enhance the mechanical performance of the membrane. Normally, the supportive layer has ˜3× of pore size as that of the selective layer and therefore has better permeability.
Another intrinsically hydrophobic potential membrane material is PP. PP membranes are strong, flexible, and compatible with a broad range of chemicals. They also have a relatively wide range of pore size and thickness that are commercially available, though for large pore sizes (>0.2 μm) the pores are no longer absolutely cylindrical and the nominal pore size is used as a representation. PP membranes have similar range of costs as PTFE membranes.
PVDF membranes are also intrinsically hydrophobic, though they have relatively higher prices as compared to other membrane materials. PVDF has been widely used in large scale industries such as membrane distillation, oil-water separation, batteries, and tissue engineering.
Although variety of hydrophobic membranes are available on the market, few of them fulfill the requirement for membrane pore size and thickness. The PVDF membranes were fabricated using a customized electrospinning setup in order to optimize membrane properties for efficient capillary-driven condensation. By the use of electrospinning, one can control the pore size and the thickness of the PVDF membrane and significantly lower down the cost of the PVDF membrane to several dollars per meter square surface area.
The wick structure is the other core component of the capillary-driven condenser design, which reduces the thermal resistance through integrating the condensate liquid film with a high-thermal conductivity structured wick of a required thickness.
Various types of structured metal wicks are available commercially. For example, metal wire cloth, perforated sheets, metal mesh, and metal foams with different thickness and porosity can be easily found on the market, as shown in Table 3. Compared to wire cloth, perforated sheets are more rigid and durable for a longer service life in harsh environments. However, both wire cloth and perforated sheets are closed-cell structures that constitute individual enclosures. On the other hand, metal foams consist of cells that are all interconnected, allowing the condensate fluid to pass through the wicking structure. The thermally conductive porous wick can include a sintered metal powder, an electrodeposited porous metal, a metal foam, a metal mesh, a laser-etched metal, a 3D printed metal, a molded surface structure, or a patterned substrate. See, for example, Wang, et al., Nature 582, pages55-59 (2020) and Richard, Bradley et al. “Loop Heat Pipe Wick Fabrication via Additive Manufacturing.” (2017), each of which is incorporated by reference in its entirety. The metal mesh can be a stack of metal meshes. See, for example, Wang, Langmuir 2021, 37, 7, 2289-2297, which is incorporated by reference in its entirety.
In addition to metal foams, sintered powder wicks are another widely used option for metal wick structures. In fact, approximately 80% of conventional heat pipes use sintered powders as wick structures. See, for example, Tang, H. et al., Appl. Energy 223, 383-400 (2018), which is incorporated by reference in its entirety. Similar to metal wicks made of metal mesh/wire cloths, sintered metal powder can provide large capillary force but relatively low liquid permeability as compared to metal foams.
The effective thermal conductivity keff of the porous metal wick is crucial to the heat transfer performance of the capillary-driven condensation. The thermal conductivity of the metal and its volumetric porosity are critical to keff. Copper has the best thermal conductivity among all the listed materials and has a mid-range price. Higher porosity of the porous wick would decrease its keff, although the wick permeability κ, which is desirable for condensate flow, increases with increasing pore size and porosity. When designing the wick structures, careful consideration needs to be made on the porosity and the pore size of the structures to ensure good balance between κ and keff.
The arrangement of the metal network could also affect keff. Highly connected metal structures like copper foams usually have higher keff than those with poorer connections even at the same porosity.
The thickness of the metal wicks determines the thickness of the condensate film and therefore plays an important role in the condensation heat transfer performance. Commercially available metal wicks such as copper foams usually have a thickness above 200 μm, limiting the heat transfer performance of the wick layer. Porous metal wicks can be fabricated through electrodeposition, such as electrodeposited copper foam. See, for example, Kim, J., et al., Electrochem. commun. 10, 1148-1151 (2008). This technique can tune the thickness of the porous metal wick through controlling the deposition time and can also be applied to large scale. The pores of the wick can have average diameters of between 5 microns and 25 microns, for example, 5 microns, 6 microns 7 microns, 8 microns, 9 microns, 10 microns, 11 microns, 12 microns, 13 microns, 14 microns, 15 microns, 16 microns, 17 microns, 18 microns, 19 microns, 20 microns, 21 microns, 22 microns, 23 microns, 24 microns, or 25 microns.
As discussed below, commercially available copper foams with the smallest thickness available on the market were used as a porous copper wick for proof of concept. Porous copper substrates were also developed with much smaller thickness based on electrodeposition.
Two commercially available materials were selected for the scalable proof-of-concept study. Porous copper foams with a thickness of ˜200 μm (thinnest that could be found on the market) and porosity of ˜70% were purchased from MTI Co. and used as the wick layer. Copper meshes with different mesh sizes (i.e., 200 mesh size, 500 mesh size, and 1500 mesh size) were purchased from TWP Inc. and Structure Probe Inc., and used as the membrane layer after hydrophobic coating. The advantages of using hydrophobized metal meshes as the membrane layer for the proof-of-concept study are (1) metal meshes are relatively easy to be bonded to the metal wick layer: (2) metal meshes with well-defined pore sizes are commercially available: (3) hydrophobic coating has well-developed deposition procedure and its lifetime (typically 1˜2 days) is long enough to experimentally validate heat transfer enhancement. Detailed properties of the two layers of materials are shown in Table 4 below.
Capillary-driven condensers were fabricated from a 1-inch diameter copper block by mechanically and then chemically polishing the end of a copper block followed by solvent and acid cleaning the surface. In order to obtain good thermal contact between the copper condenser block, the copper foam wick, and the copper mesh membrane, the copper foam was diffusion-bonded and the copper mesh to the copper block using a furnace.
A schematic and an image of the latest diffusion bonding assembly are shown in
Three different hierarchical copper surfaces were designed and fabricated through diffusion bonding, as shown in the scanning electron microscopy (SEM) images in
The permeability of the wick layer of the hierarchical copper samples was measured, which is the 200 μm-thick copper foam, under various processing conditions, such as different pressure for the diffusion bonding/hot press procedure and different heating time for the sample to be kept in the high temperature furnace. The permeability measurement was conducted by following an established procedure described in the literature (see, for example, reston, D. J. et al. Gravitationally-Driven Wicking for Enhanced Condensation Heat Transfer. (2018), which is incorporated by reference in its entirety).
For the proof of concept study, the membrane layer was fabricated by hydrophobizing the copper mesh. The hydrophobic coating was achieved vapor deposition using FAS ((Heptadecafluoro-1,1,2,2-tetrahydrodecyl) trimethoxy silane) to form conformal hydrophobic coating on metal surfaces. This specific coating exhibits a consistent water contact angle of ˜105°.
The details of the coating procedure are as follows. First, the copper sample was cleaned with acetone, ethanol, isopropanol, and water in sequence to remove potential hydrocarbon contaminants. Then, the sample was dipped in to an HCL solution (0.2 M) for 30 seconds to remove oxides. After the samples are rinsed with water and dried with nitrogen, the sample was oxygen plasma cleaned for 20 minutes. With bombardment of air plasma, the copper surface would be more active and easier to form a conformal coating.
Then, the sample was placed into a sealed bottle. Along with the porous copper samples, 800 μL of FAS in toluene solution (5 V %) was also put into a small beaker, which sat beside the samples to be coated. The sealed bottle was placed in an oven at 100° C. for 3 hours, during which the FAS was coated onto the porous copper sample via vapor phase deposition. Finally, the samples were taken out of the furnace and cooled down to room temperature in the fume hood. A schematic of the coating process in illustrated in
Ideally, only the copper mesh layer needs to be hydrophobized for the proof-of-concept study, and the copper foam layer should remain hydrophilic to promote condensation nucleation. Biphilic coating can be made by depositing hydrophobic coating when covering the bottom-layer structures with a protective coating which can be washed away easily after the hydrophobic coating. For this study, biphilic coating was applied to the hierarchical copper sample by utilizing a heat curing photoresist (AZ3312) as the protective layer for the copper foam structures. The detailed procedure includes following steps. First, the sample was cleaned with acetone, ethanol, isopropanol, and water, followed by 10 minutes of argon plasma cleaning. Then, the sample was dipped into a beaker of photoresist AZ3312 and allowed the photoresist to climb up into the porous copper under capillary force. Once all the copper foam area was covered by photoresist, the sample was baked in an oven at 100° C. for 10 minutes until the photoresist was solidified inside the porous copper sample. Next, 20 minutes of oxygen plasma was applied to remove the top layer photoresist off the hierarchical sample, exposing the hydrophilic copper mesh to be hydrophobized. The photoresist-protected hierarchical copper sample was then deposited with FAS coating using the procedure illustrated in
Although the permeability of the copper foam (1E-11 m2) was high enough to prevent flooding at low surface subcools, flooding could potentially occur at local defects such as broken mesh pores, which was hardly evitable during the fabrication process. In order to further facilitate condensate drainage, microchannels were machined on the copper foam layer, as shown in
An analytical model was developed based on 1D conduction through liquid-filled porous copper to predict the heat transfer performance of the three hierarchical sample surfaces. To calculate the effective thermal conductivity of a copper mesh, the following equation was adopted, which is developed and validated by previous literature (see, for example, Reay, D. A., et al., Heat Pipes (Sixth Edition) 65-94 (Butterworth Heinemann, 2014), which is incorporated by reference in its entirety):
where ε is the volume faction of the solid phase. ks is the thermal conductivity of the solid (copper), and kl is the thermal conductivity of the liquid (water).
The effective thermal conductivity of the wick layer (copper foam) can be calculated using the volumetric average value as been shown in previous literature (see, for example, Preston, D. J., et al., Langmuir 2018, 34, 15, 4658-4664, which is incorporated by reference in its entirety):
Heat flux and fluid flow were analytically solved under an environmental condition with vapor temperature of 35° C. and surface subcool up to 5° C. This vapor condition is normally seen in steam power plant condensers and this condition was the goal to achieve during the experiments. Conventional filmwise condensation heat transfer predicted by the Nusselt model is used as the benchmark.
Heat transfer performance of the fabricated sample surfaces was experimentally characterized in a controlled environmental chamber, as shown in
A Pirani gauge was installed on the chamber to accurately monitor the chamber pressure under pumping process, as shown in
Condensation heat transfer on a flat copper surface was measured and compared to the Nusselt model prediction to validate the capability of the environmental chamber.
For the proof of concept study, three hierarchical copper samples were tested, all of which were based on 1500-mesh-size copper mesh covered copper foam structures. Samples with different wettability (i.e., fully hydrophobic and biphilic) were tested to investigate the influence of surface wettability on the condensation heat transfer performance.
The 50% enhancement measured on the biphilic hierarchical copper sample was promising, although it was lower than what was expected from the model prediction. The following reasons are the hypothesis for the discrepancy between the modelling results and the experimental results. (1) The micro-channels decreased the effective thermal conductivity of the copper foam layer, which was not captured in the model: (2) The copper foam had a poor connection in its vertical direction as compared to its horizontal direction, but the vertical direction was critical for the heat transfer experiment. Therefore, the effective thermal conductivity of the copper foam was overestimated by calculating a volumetric average value. (3) There were inevitable defects such as broken pores in the copper mesh layer due to the diffusion bonding process. These defects would affect flooding criteria and heat transfer performance. Those defects were not considered in the model. Thermal conductivity of the micro-channelled hierarchical copper sample needs to be measured or estimated using a more accurate model. The design of the microchannels (channel width, channel height, distance between nonboring channels) can also be optimized. Combining a thinner copper wick layer with an optimized microchannel structure is expected to further enhance the heat transfer performance of the capillary-driven condenser.
As mentioned earlier, porous metal wick can be fabricated in a scalable manner by electrodeposition. A recipe to electrodeposit a single layer of inverse opal copper was developed. See, for example, Carlos, D. D., et al., Langmuir 2021, 37, 43, 12568-12576, which is incorporated by reference in its entirety. By using templating spheres with 10 μm diameter, a monolayer inverse opal copper was fabricated with roughly 65% porosity, 5 μm thickness, and a model-predicted porosity of 5E-11 m2 permeability. The remaining challenge in the fabrication of inverse opal copper is sustaining a robust attachment and close to full coverage of the templating spheres during the electrodeposition step, which can be critical to the scale of the surface area to be fabricated.
Sintering copper powders is another scalable way to fabricate porous copper. Sintering copper powders using three different powders was attempted: spherical powders with diameter <10 μm, spherical powders with diameter <50 μm, and dendritic powders with size <45 μm. The resulting porous copper structures are shown in the SEM images in
Due to the size of the tube furnace, sintered copper powder surfaces were fabricated on the scale of 1 cm. However, sintering copper powers is a highly scalable technique as can be seen in the heat pipe industry. For the capillary-driven condenser, a small thickness (ideally under 200 μm) is required for the porous wick layer in order to reduce thermal resistance of the wick layer. Microporous copper wick monolayers have been achieved by sintering in the literature (see, Hoenig, S. H. & Iii, R. W. B. Journal of Heat Transfer 140. 1-7 (2018), which is incorporated by reference in its entirety), although depositing a monolayer of copper powders over a large surface area is still challenging. Sintering copper powders with smaller size could keep the wick thickness small while depositing multiple layers of powders. However, the wick permeability would decrease with the size of the powders. The permeability of a sintered powder wick can be estimated by:
where κ is the permeability of the sintered powder wick, dpowder is the size/diameter of the powders, and ϕw is the porosity of the sintered powder wick. Sintered copper powders has shown a porosity of ˜50% in literature. See, for example, Leong, K. C., et al., Journal of Porous Materials, 4, 303-308 (1997), which is incorporated by reference in its entirety. Sintering copper powders with diameter of ˜50 μm would be good to balance both the thickness and the permeability of the wick layer. Assuming a sintered copper wick made by four layers of 50 μm copper spheres with a porosity of 50%, the copper wick's permeability was expected to be ˜1.7E-11 m2, which is comparable to the permeability of the copper foam used for the proof of concept study.
Electrospun membranes are described below.
A heat and mass transfer model for electrospun-fiber covered porous copper based on vapor transport through the membrane pores and heat conduction through the condensate filled porous copper was developed.
(a) is commercially available. (b) is a structure fabricated in lab. (c) is a common structure in heat pipe industries and can be fabricate. All three plots shown in
The lines in
Overall, the model guided fabrication of a membrane that is thin (on the order of 10 μm), highly porous (90%), and has a large pore size (on the order of 1 μm); it was also important to make sure that flooding would not occur on this membrane under a sufficient subcool (at least 5° C.). Once the parametric optimization for electrospinning fibrous membranes is completed, it will be possible to combine an optimized membrane with a porous structured copper substrate to make a scalable version of the capillary-driven condenser. The structure is expected to show over 5x enhancement on the electrospun membrane covered sintered copper powder without flooding under a subcool of 5° C., as predicted by model.
Electrospinning is an easily tunable technique that utilizes the force balance between the electrostatic force and the solution surface tension to fabricate nanofibrous membrane, is promising for a wide range of applications. Morphology of the electrospun membrane can be directly related to its performance. During an electrospinning process, many factors can affect the morphology of the products, such as properties of the solution, voltage supply, needle distance, feeding rate of the solution, and the electrospinning time. Effective parametric study can be used to optimize the fabrication of electrospun membrane such that the ideal morphology of the electrospun membrane can be achieved by optimizing the most important parameters. Poly(vinylidene fluoride-co-hexafluoropropylene) was chosen as an example for the current study due to its interesting and versatile functionalities such as superior hydrophobicity, high free volume, and piezoelectricity.
As noted above, thermoelectric power generators, which typically use a Rankine cycle, provide the majority of the electricity produced in the US. See, for example, ICF-International, Catalog of CHP Technologies, in Combined Heat and Power Partnership Program, E.P.A. (EPA), Editor. 2008, U. S. Environmental Protection Agency (CHPPP): Washington, D.C., which is incorporated by reference in its entirety. However, they also withdraw the largest amount of water from US water bodies in order to condense the steam generated from the power plant. There is significant interest in enhancing the efficiency of condenser designs, which will not only improve power production but also decrease the amount for water needed for condensation. Significant efforts have focused on advancing condenser designs with higher performance for steam power plants. Various condenser designs have been proposed and incorporated in existing power plants. See, for example, Hao, M., et al., Program on Technology Innovation: New Concepts of Water Conservation Cooling and Water Treatment Technologies. 2012, Electric Power Research Institute (EPRI): Palo Alto, which is incorporated by reference in its entirety. These condensers, however, typically rely on filmwise condensation (
Filmwise condensation is not desired due to the large thermal resistance to heat transfer. See, for example, Nusselt, W., The surface condensation of water vapour. Zeitschrift Des Vereines Deutscher Ingenieure, 1916. 60: p. 541-546, which is incorporated by reference in its entirety. Over the past eight decades, dropwise condensation, where droplets roll off at sizes approaching the capillary length and clear the surface for re-nucleation (
As described herein, capillary-driven condensation (CDC) can involve a structure in which a porous hydrophobic membrane atop a wicking structure on a condensing surface drives liquid transport and removal from the surface via a capillary pressure gradient along the wicking surface towards an exit port, as shown in
The invention disclosed herein relates to a robust new approach to enhance condensation heat transfer for steam power plants via capillary-driven condensation (
The key advantages of this novel mode of condensation are: i) a decrease in the thermal resistance of the condensation process due to the increased effective thermal conductivity of the liquid-wick layer; and ii) capillary-assisted removal of the condensate film which provides orders of magnitude higher driving force than gravity. Equally important, this condensation mode promises to achieve both an enhanced heat transfer coefficient and overall heat flux in a robust manner, i.e., without concerns of degradation of surface coatings typical for promotion of dropwise condensation.
Detailed and systematic studies have supported that this represents a promising solution for steam power plant condensers. Towards this goal of enhancing the heat transfer coefficient and ensuring superior robustness. Some of these studies described herein include:
Develop porous membranes and wicking structures for capillary-driven condensation. Various wicking structures and porous hydrophobic membranes have been designed to reduce the thermal resistance and enhance capillary-driven flow.
Experimentally investigate capillary-driven condensation on flat substrates and tube substrates. Condensation heat transfer performance has been experimentally studied and compared with traditional filmwise condensation on various samples.
Optimize the capillary-driven condensation structure with model development. A physics-based model has been developed to predict and optimize condensation heat transfer, and validated by experiments.
Incorporate capillary-driven condensation structure to demonstrate scaled-up proof-of-concept operation. Experiments performed on tube bundles in industrially relevant conditions can provide additional insights.
Based on the success of these objectives, a robust condenser design can be implemented for steam power plants with greater than five times enhancement in heat transfer coefficients compared to conventional filmwise condensation. Due to the improved heat transfer coefficient of condensation the steam condensation temperature, and accordingly the turbine back-pressure, can be reduced by up 4° C. and 0.7 kPa, respectively. As a result, the overall heat rate of a typical power plant can be expected to decrease by 1.5%, leading to an additional 13.80 MW of generated power for a 950 MW plant and a commensurate savings in water withdrawal and usage. Implementation of this method has a simple payback period of roughly one year.
In capillary-driven condensation, condensation occurs on the hydrophilic micro/nanostructured wick, and the condensate is then forced out due to the capillary pressure buildup at the menisci formed in the porous hydrophobic membrane. The presence of the structures and the resulting capillarity helps maintain a stable liquid film while driving liquid flow. By tailoring the size of the pores in the membrane and the geometry of the wicking structure, the capillary pressure generated can be maximized and the flow rate of the condensate can be optimized to increase the rate of condensation that the wicking structure can support. This approach promises to be completely passive, robust, and can harness an order of magnitude higher heat transfer coefficients than conventional filmwise condensation. Such a strategy can be a development breakthrough for developing a condensation strategy for power plants.
The capillary-driven condensation mode can be modelled and compared to traditional filmwise condensation theory. This model and the predicted enhancement over filmwise condensation motivate the need to study and further development of the concept of capillary-driven condensation for applications in steam cycle power plants.
Filmwise condensation can be described as follows. Typically, condensation on industrial materials such as metals results in a fluid film which sheds due to the gravitational force. This hinders heat transfer by adding a significant conduction thermal resistance. Nusselt modeled this phenomenon using lubrication theory, and the model has been shown to agree very well with experimental data. The heat transfer coefficient determined by this model for a horizontal tube is:
Capillarity can also be important. On a high energy surface, interaction between the liquid and the solid surface leads to the formation of a liquid-vapor interface called the meniscus. The surface tension of the interface leads to a pressure difference across the meniscus. This pressure difference is the capillary pressure Pcap and can be described by the Young-Laplace equation:
Enhancing fluid flow with a capillary pressure gradient can be an important feature of the structures described herein. In the design described herein, capillary transport through a thermally conductive wick is driven by capillary pressure generated by the convex interfaces formed with a hydrophobic membrane encasing the wicking materials. This design removes condensate more rapidly than filmwise condensation on a bare (without a wick) surface, thereby enhancing heat transfer. A design is shown in 21B as compared to traditional filmwise condensation in
The performance of this design is determined using Darcy's law for porous media to model the flow within the wick. The mass flux of condensate into the wick, {dot over (m)}″, which is proportional to the heat flux by the constant factor of the latent heat of the fluid, is assumed spatially uniform due to the wick's uniform thickness (twick) and thermal conductivity (keff) as detailed in
The previous two equations (equations (24) and (25)) relate the heat flux with the required pressure drop. Therefore, for a given maximum capillary pressure drop, one can determine the maximum achievable heat flux which can occur before the capillary pressure generated by the porous membrane can remove all of the condensate.
The results of the modeling effort are shown in
These results guide the fabrication and synthesis process of the wicking surface, where high permeability wicks are desirable. A key takeaway from the model results in
The potential economic benefit of capillary-driven condensation for a typical 950 MW nuclear fired power plant (heat rate of 2.942 KW/KW at 3 in Hg abs) was analyzed following the analysis in Webb, R., Enhanced Condenser Tube Designs Improve Plant Performance. Power Magazine, 2010, which is incorporated by reference in its entirety. Estimated costs of the structured porous wick and hydrophobic porous membrane are shown in Table 5 based off of necessary materials to modify a condenser with 23,150 tubes with an outer diameter of 28.6 mm and 13.4 m long. A porous aluminum wick 5 mm thick and a 30 μm thick Nafion membrane were used for cost estimation.
An alternative system can be based on a porous copper powder wick (0.2 mm thick) and PVDF membrane (pore size ˜1 μm) (Alternative membrane materials: PTFE, PP) are shown in Table 6.
With the analysis shown in
The performance of capillary-driven condensation can depend on a number of factors. One factor includes wicking structures and porous hydrophobic wicks. The wick can be a structured wick. The wick can serve at least two main roles: (1) it can serve as a structural support to hold the hydrophobic membrane away from the surface in order to allow condensate to flow underneath it; and (2) it can enhance the effective thermal conductivity of the condensate layer that is flowing beneath the hydrophobic membrane by providing a path of lower resistance for heat transfer through the wick itself. In certain circumstances, a wick can have high permeability and high thermal conductivity, which are two of the most important parameters for enhancing condensation heat transfer. High permeability is required to reduce viscous flow losses: meanwhile, high thermal conductivity decreases the effective thermal resistance of the condensate-wick layer. The effect of an increased permeability is demonstrated in
The other critical aspect of the capillary-driven condensation design is the porous hydrophobic wick. This membrane is responsible for holding the condensate inside of the structured wick and also generating the capillary pressure (according to Equation (23)) that causes flow of the condensate through the wick towards the exit ports. Several important parameters of the membrane are the pore radius, the thickness, and the intrinsic contact angle. The effect of one of these parameters, the pore radius, is illustrated in
In addition to the design of both the wick and the hydrophobic membrane separately, another consideration is the interfacing of the two when the capillary-driven condenser is assembled. Both the wick and the membrane can be wrapped around a typical tube condenser during fabrication in a scalable and cost-effective manner as discussed below, the approach described herein can be a robust and long-term solution for improved condensation heat transfer in steam power plants. Approaches can include interfacing or bonding the membrane directly onto the wicking material before application to condenser tubes, such as physical, thermal, or stress-based attachment.
Wicking enhanced filmwise condensation can be characterized experimentally. To investigate capillary-driven condensation heat transfer performance with water as the working fluid, a custom built multi-purpose environmental chamber can be used as shown in
By monitoring cooling power, surface temperatures, ambient conditions (including presence of non-condensable gases, pressure and temperature, etc.), the performance of the condensation mode can be analyzed: Qcooling=U·Asurface·ΔTLMITD Where ΔTLMITD is the log-mean temperature difference between the surface temperature and the vapor environment temperature and Qcooling is the cooling power which is dependent on the mode of cooling: for a chiller: Qcooling={dot over (m)}·cp·(Tout−Tin) where m is the mass flowrate of the refrigerant, cp is its specific heat, and Tin and Tout are the inlet and outlet refrigerant temperatures. Qcooling for a Peltier device would be: Qcooling.Peltier=P·I where P is the Peltier coefficient and I is the current drawn by the device. The overall heat transfer coefficient U can be used to determine the condensation heat transfer coefficient as a function of the cooling power Qcooling and the condensation temperature difference.
Apart from the operating fluid, various combinations of wick materials and geometries will also be tested. In situ visualization of the condensation process (similar to
Prior work includes a study on condensation enhancement for hydrocarbon liquids using only a gravitational pressure gradient in a wick (no hydrophobic porous membrane). The results included a 3× enhancement in heat transfer performance, as shown in
An accurate numerical model can be developed which captures the liquid-vapor interface and its effect on the fluid transport to better quantify the capillary-driven condensation heat transfer performance on different wick structure geometries. The numerical model can allow for the design of wick structures to maximize the heat transfer enhancement. A finite volume model based on CFD simulations was developed to predict the condensate liquid flow rate for different wick structure geometries. This model will allow one to understand the maximum heat flux that can be achieved via this mode of condensation (i.e., the maximum heat flux before liquid bursts through the hydrophobic top surface). The model also allows us to predict the heat transfer coefficient for different wick structure geometries for comparison with filmwise condensation.
A modeling framework that was developed for thin-film evaporation on micropillar array wick surfaces, which is a liquid-to-vapor phase change process based on similar design and physics, was modified. See, for example, Zhu, Y., et al., Langmuir, 2016. 32(7): p. 1920-1927, which is incorporated by reference in its entirety. The micropillar geometry can beneficial for systematic understanding of experiment where well-defined pillar arrays can easily be fabricated. The modified model for capillary-driven condensation can predict liquid velocity, pressure, and meniscus curvature along the wicking direction which is achieved by conservation of mass, momentum and energy. Specifically, models can accurately predict the three-dimensional meniscus shape, which varies along the wicking direction with the local liquid pressure (
Initial calculations based on the preliminary modeling framework were very promising heat transfer performance in this mode of condensation.
The knowledge gained from capillary-driven structure development, heat transfer characterization experiments, and numerical modeling described above can guide the design of optimized capillary-driven condensation structures. Scaled-up condenser prototypes for the proof-of-concept on large tubes (˜10-100 cm length) in tube bundles can be fabricated and tested using a Low Pressure Condensation Unit (LPCU).
A porous hydrophobic membrane atop a wicking structure on the condensing surface drives liquid transport and removal from the surface via a capillary pressure gradient along the wicking surface towards an exit port. Based on the success of the objectives described, a robust condenser design for steam power plants with >5× enhancement in heat transfer coefficients compared to conventional filmwise condensation was developed. Due to the improved heat transfer coefficient of condensation the steam condensation temperature, and accordingly the turbine back-pressure, can be reduced by up to 4° C. and 0.7 kPa, respectively. As a result, the overall heat rate of a typical power plant can be expected to decrease by 1.5%, leading to an additional 13.80 MW of generated power for a 950 MW plant and a commensurate savings in water withdrawal and usage. Implementation of this method has a simple payback period of roughly one year.
As observed in the condensation experiment with the hierarchical copper sample, microchannels can help with fast drainage of the condensate especially in the case where the hydrophobic membrane has defective pores which could cause local flooding/bursting out of the membrane. A few considerations should be taken when adding microchannels to the capillary-driven condenser design. First and foremost, microchannels will decrease the effective thermal conductivity of the metal wick layer as it would break the continuous connection of metal network for the efficient thermal transport. Therefore, the addition of microchannels should be just enough to prevent the membrane from flooding. When designing the geometry of the microchannels, the viscous pressure loss associated with the process of the condensed water traveling through the porous wick and finally exiting through the microchannel (route is shown as the arrow in
Taking one microchannel and its neighboring metal wick block as a unit of interest. The viscous pressure loss occurred inside the porous metal wick can be described by 1D Darcy's Law:
According to the mass conservation of the condensed water, {dot over (V)}(x) can be expressed as a function of the condensation heat flux q (W/m2) and the latent heat of water hfg as
For laminar flow (Re>2300), the viscous pressure drop occurred inside a rectangular microchannel can be calculated by
For turbulent flow (Re>2300) in a rectangular channel, the Fanning friction factor can be calculated by the Blasius equation:
where Re* is the corrected Reynolds number for rectangular channel geometries
See, for example, Adams, T. M., Abdel-Khalik, S. I., Jeter, M., Qureshi, Z. H., 1997. An experimental investigation of single-phase forced convection in microchannels. Int. J. Heat Mass Transfer 41 (6-7), 851-857 and Jones Jr., O. C., 1976. An improvement in the calculation of turbulent friction in rectangular ducts. J. Fluids Eng. 98, 173181, each of which is incorporated by reference in its entirety. In the preferable case where the condensed water exits through the microchannel (as shown in the arrow in
Comparing the viscous pressure loss calculated by Equation (26) and Equation (28) would determine if the addition of the microchannels can help with the condensate drainage. An optimized microchannel structure should ensure that it accelerates condensate drainage such that flooding would not occur at the defected membrane pores: meanwhile, the microchannel's negative impact on the effective thermal conductivity of the metal wick layer should be minimized.
A microchannel wick can be designed with microchannels with varying geometries. There can be no foam or other structure, just microchannels. The microchannels can have a height of less than 30 microns, a channel wall of less than 100 microns, and a channel pitch of less than 1 mm.
In an effort to fabricate a scalable, robust, and cost-effective capillary-driven condenser, intrinsic hydrophobic membranes were fabricated based on electrospinning. Electrospinning is a sophisticated technique for fabricating nano- and microfiber membranes, which has been applied to various industries such as water desalination, oil separation, and water harvesting. Electrospinning to directly deposit fibers atop porous metal wick to fabricate capillary-driven condensers can be used in a scalable way.
A customized electrospinning set up was built, as shown in
An aluminium foil covered copper plate is used as the fiber collector and is connected to a ground wire. The high voltage difference across the needle and the collector drives the polymer solution to spin onto the collector in a form of nano- or microfibers, under the counteractions of electrostatic force and surface tension experienced by the solution. The electrospinning process involves many operation parameters, some of which are critical to the fiber/membrane fabricated by the process. Some key parameters include voltage being applied, distance between the needle and the collector, the solution feeding rate, the properties of the polymer solution, and the time the electrospinning process lasts. Without limitation, electrospinning PVDF-HFP was initially studied since this polymer is intrinsically hydrophobic and is known to form uniform membrane pores.
Three properties of the membrane are critical to vapor transport and heat transfer through the membrane layer: pore size, thickness, and porosity. For electrospun fibrous membranes, the pore size can be proportional to the fiber diameter. In order to characterize these important properties, several techniques were used as detailed below.
Scanning electron microscopy (SEM) images were taken to help us characterize the morphology of the electrospun fibers, as shown in
The thickness of the membrane was measured by a micrometer with a resolution of 1 μm. The average of three independent thickness measurements was calculated at the thickest part of the electrospun membrane and used it as the membrane thickness. In addition, the membrane weight and the membrane surface area were measured to calculate membrane porosity. A consistent porosity of ˜80% was obtained through all of the PVDF-HFP samples, which is in good agreement with the literature. See, for example, Ahmed, F. E., et al., DES 356, 15-30 (2015), which is incorporated by reference in its entirety.
An optimized parametric study on the electrospinning process based on fractional factorial design (FFD) was completed. Investigating the effects of each independent parameter (full factorial design) tends to be laborious and redundant. In statistics, fractional factorial design (FFD) is a powerful tool for reducing the experimental burden while recognizing the major effects. FFD consists of a carefully chosen subset (fraction) of the experimental runs of a full factorial design, identifying the most important parameter to the response of interest. The effects of multiple fabrication parameters on the morphology of the electrospun membrane can be studied efficiently by FFD.
The FFD method used in this study varied factors together and fitted the results using a polynomial method. It could determine how factors interact but could have errors due to polynomial assumption. Therefore, FFD was used for screening parameters before moving on to full factorial study on the most important parameter.
During the screening study, the effects of the three important fabrication parameters were studied, namely the voltage supply, the discharge/needle distance, and the syringe pump feeding rate on the morphology of the electrospun membrane. Table 8 shows the selected key parameters studied and their variation range. The variation range for each parameter was empirically determined to ensure successful electrospinning without bead formation. The duration of the electrospinning was held constant (60 minutes) and the solution concentration unchanged (15% wt PVDF-HFP in 3:7 vol ratio of N,N-dimethylacetamide/acetone) throughout all runs of experiments. The environmental temperature and humidity was controlled at around 23° C. and 30% respectively. The electrospun fiber was left in a well ventilated area overnight for complete evaporation of the solvent before SEM imaging.
Table 9 shows the subset of the experimental runs guided by the fractional factorial design (solution concentration 15 wt % and electrospinning time of 60 minutes). The fractional factorial design simplified a full 8-run parametric study into 4 runs. Each set of parameters were run though at least 2 times under similar environmental conditions and the morphology of the resulted membranes was characterized with SEM and a capillary flow porometer.
To optimize the fabrication of the electrospun PVDF-HFP membrane for the best performance of the capillary-driven condensation, membrane pore size needs to be optimized. A larger membrane pore will allow easier vaport transport, but the pore size should not exceed the maximum limit of the membrane pore size at which the capillary pressure cannot hold the condensed water in place and cause flooding. FFD tells us that the solution feeding rate is the most important parameter among the three key paratemers that were studied (voltage supply, needle-to-collector distance, and feeding rate). Future work entails a detailed full factorial study on the effect of solution feeding rate on the electrospun membrane pore size. The membrane thickness was not studied here because membrane thickness is a very tunable characteristic by simply increasing the duration of electrospinning to increase membrane thickness. Moreover, the effects of the duration of electrospinning on the fiber diameter were studied and demonstrated that there is little correlation between the duration of the electrospinning time and the resulted fiber diameter, as shown in
Effects of solution aging on fiber diameter was studied.
Getting a strong bonding between the electrospun membranes and the porous metal substrates can be challenging. In a preliminary test, electrospun membranes directly deposited on untreated substrates could be peeled off the substrates easily. However, it was found that solvent clean treatment and plasma clean treatment on the substrates could enhance the attachment between the electrospun fibers and the metal substrates. For this preliminary test, copper pillar surfaces were used as the porous metal substrate for easier imaging. These are square-shape copper pillars with 100 μm height, 100 μm neighborhood distance and 100 μm side length on a 3 mm thick copper substrate. The substrates were cleaned with acetone, ethanol, isopropanol and water, dried the substrates with nitrogen flow, and then plasma cleaned the substrates with oxygen plasma for 10 minutes. These samples were used for electrospinning right after the cleaning process. The resulting electrospun fibers showed larger preference to attach to the metal surfaces than to the fiber themselves, as shown in
Heat treatment was found to be an effective approach to enhance the bonding between the electrospun membrane and the porous metal substrate. During a preliminary study, the porous copper substrate was put on a grounded hot plate and during the first 5 minutes of electrospinning heated the substrate to 160° C., which is slightly higher than the melting point (143° C.) of the polymer used for electrospinning (PVDF-HFP). By heating up the substrate, it was expected the electrospun fibers deposited on the substrate during the first few minutes to melt and form better and more uniform bonding with the porous metal substrate, after which the electrospun fiber could form a strong bonding to the substrate driven by the attraction between PVDF-HFP fibers without the need of the heat. It was also found that the geometry of the metal wick has a significant effect on the bonding between the electrospun membrane and the metal wick. Electrospun PVDF-HFP membranes were formed on two different substrates with the same heat treatment (kept of a 160° C. hot plate during the first 5 minutes of electrospinning). Better bonding was formed on the 200 μm pore size copper foam than the membrane spun on the 100×100 size copper mesh (200 μm equivalent pore size).
Based on the above preliminary experiments, the following strategy for fabricating the capillary-driven condenser tube in a scalable way can be followed:
Fabricating the porous metal wick layer by template-assisted electrodeposition or sintering. The type of the metal being used in this step is not fixed by should be compatible with the tube material. Conventional materials being used in condensers include copper, copper-nickel alloy, stainless steel, etc.
Machining of exit ports or channels on the porous copper wick layer. This step can be integrated into the first step or post-processed after the first step. Depending on the geometry for the channels, different tools can be used.
Adding a layer of hydrophobic membrane on top of the porous metal wick. It is believed electrospinning has a high potential to customize the hydrophobic membrane in the current application. Securing the porous wick covered condenser tube on a rotating system (similar to the electrospinning drum collector) would allow the electrospun fiber to be uniformly deposited onto the tube surface. On a tube geometry, the bonding of the electrospun membrane would be secured by the alignment of the nanofibers. Heating the tube substrate above the melting point of the polymer being electrospun during the first few minutes of electrospinning can further enhance the bonding between the membrane and the porous metal wick layer.
Machining of exit ports or channels on the membrane layer. This can be done by simply locally cutting or melting the membrane.
Mechanical clamping can be added to further secure the membrane in place.
This equation shows the competition between the capillary pressure Pcap which is a function of the pore diameter dp against the viscous pressure drop in the wick ΔPwick which is a function of the fluid transport length Lf, and the wick geometry. Note that the viscous pressure drop is proportional to the square of the fluid transport length in a simple model utilizing the Darcy equation for flow in porous media. Therefore, flooding is strongly dependent on this term. In the case of a micropillar wick, the membrane geometry is characterized by h and d/l, which are the thickness of the wick (pillar height) and the pillar diameter to pillar pitch ratio, respectively, whose quantities determine the pore size of the wick, and its porosity.
If elasticity is utilized, a membrane can wrap around the tube and be held by tension without the need to have adhesives for attachment. Moreover, mechanical clamping can be utilized in some configurations to avoid the necessity to utilize bonding methods which can require adhesives or multistep coatings to achieve stronger bonding. For example, the membrane sections can be held at the bottom by stitching, mechanical clamping or other method that does not require adhesives. By eliminating the requirement to bond a membrane to the wicking structure in the conventional way of using adhesives, the membrane only needs enough clamping force to sustain the required capillary pressures in its membrane pores during operation. All materials in this category should be stable in a steam environment.
Modelling predicts ˜4× to 12× heat transfer enhancements compared to filmwise condensation. By comparison, dropwise condensation is known to have ˜5×-˜7× enhancements measured experimentally. The pore size can range from about 1-20 microns. Commercial membranes are typically thicker there will be a larger vapor transport resistance. However, increasing the pore size also allows better vapor transport. This can mean less capillary pressure to sustain fluid flow in the wick.
Thus, the model can predict enhancements beyond those experimentally measured for dropwise condensation.
Each of the following references, noted above, is incorporated by reference in its entirety.
The following United States patent documents as also incorporated by reference in its entirety:
It should be understood that the subject matter defined in the appended claims is not necessarily limited to the specific implementations described above. The specific implementations described above are disclosed as examples only.
This application claims priority to U.S. Provisional Patent Application No. 63/340,799, filed May 11, 2022, U.S. Provisional Patent Application No. 63/189,555, filed May 17, 2021, each of which is incorporated by reference in its entirety.
This invention was made with government support under grant number DE-FE0031677 awarded by the Department of Energy. The government has certain rights in this invention.
Filing Document | Filing Date | Country | Kind |
---|---|---|---|
PCT/US2022/029713 | 5/17/2022 | WO |
Number | Date | Country | |
---|---|---|---|
63340799 | May 2022 | US | |
63189555 | May 2021 | US |