Universidad de Castilla-La Mancha FACULTAD DE
Transcription
Universidad de Castilla-La Mancha FACULTAD DE
Universidad de Castilla-La Mancha FACULTAD DE CIENCIAS Y TECNOLOGÍAS QUÍMICAS DEPARTAMENTO DE INGENIERÍA QUÍMICA TESIS DOCTORAL: VALORIZATION OF POLYSTYRENE WASTES USING NATURAL TERPENES AND HIGH PRESSURE CO2 Presentada por Cristina Gutiérrez Muñoz para optar al grado de doctor por la Universidad de Castilla-La Mancha Dirigida por: D. Juan Francisco Rodríguez Romero Dña. Mª Teresa García González D. Juan Francisco Rodríguez Romero Catedrático de Universidad Departamento de Ingeniería Química Universidad de Castilla-La Mancha y Dña. Mª Teresa García González Profesor titular de Universidad Departamento de Ingeniería Química Universidad de Castilla-La Mancha Certifican que: Cristina Gutiérrez Muñoz ha realizado bajo su dirección el trabajo titulado “Valorization of Polystyrene wastes using natural terpenes and high pressure CO2”, en el Departamento de Ingeniería Química de la Facultad de Ciencias y Tecnologías Químicas de la Universidad de Castilla-La Mancha. Considerando que dicho trabajo reúne los requisitos para ser presentado como Tesis Doctoral expresan su conformidad con dicha presentación. Ciudad Real, a ____ de ____________ de 2014 Juan Francisco Rodríguez Romero Mª Teresa García González Resumen El volumen de plásticos producidos en Europa crece anualmente, debido a su bajo precio, la versatilidad que presentan y el amplio campo de aplicaciones en los que se incorpora. Sin embargo, tras su uso, se convierten en residuos que es necesario reciclar para proteger y conservar el medioambiente, tal y como marcan las políticas Europeas de gestión de residuos, que se centran en la reducción de los impactos negativos que producen los plásticos. Afortunadamente, a lo largo de la última década se está produciendo una tendencia positiva en la gestión de los residuos plásticos, hacia su recuperación y reciclado. De este modo, el reciclaje de residuos plásticos se presenta como una alternativa medioambientalmente ventajosa, pero que debe tener en cuenta posibles beneficios económicos para considerarla como óptima. La recuperación de residuos plásticos en forma de productos de alto valor añadido, hace posible la viabilidad de los procesos de reciclaje por las siguientes razones: reducción del consumo de materias primas, recursos y energía, ahorro de espacio en vertederos, beneficios medioambientales y sociales. El presente trabajo de investigación, se centra en el reciclaje de residuos de Poliestireno (PS) mediante una tecnología medioambientalmente sostenible, que permita la obtención de un producto de alto valor añadido. Se seleccionó el Poliestireno por tratarse de un plástico no biodegradable, que consume inmensas cantidades de energía y recursos durante su producción y que debido a su versatilidad y su escasa vida útil, termina depositado en vertederos o desintegrado en pequeños fragmentos en el océano. Con el fin de favorecer el reciclaje de residuos de Poliestireno, se tienen que diseñar nuevos procesos que permitan reducir el impacto medioambiental, los costes implicados y que a su vez, mejore las propiedades de los productos recuperados respecto a los originales o a los obtenidos mediante las tecnologías tradicionales (reciclado mecánico, reciclado químico o valorización energética). De acuerdo con esto, y teniendo en cuenta los principios de la Química Verde, se propone un nuevo proceso. Éste consiste en la disolución de los residuos de Poliestireno en disolventes naturales de la familia de los terpenos, con lo que se logra reducir el volumen de los residuos (disminuyendo los costes de transporte), eliminar impurezas insolubles en los aceites terpénicos, y disolver selectivamente el polímero a partir de una corriente residual que contenga mezcla de distintos productos plásticos. Además, el polímero permanece intacto, ya que el disolvente no modifica su estructura interna, ni favorece su degradación. A continuación, es necesario separar el disolvente del polímero, para ello, se adiciona un antisolvente que permite la precipitación del polímero y la eliminación del disolvente. El antisolvente elegido en este trabajo, es el dióxido de carbono (CO2) en condiciones de alta presión, ya que los aceites terpénicos presentan una elevada solubilidad en él, y su adsorción sobre el polímero facilita el procesado. Además, el CO2 puede utilizarse como agente espumante, por lo que incorporado al proceso, da lugar a espumas microcelulares, que presentan un alto valor añadido. Cabe destacar que a lo largo del proceso descrito, no se producen corrientes residuales, ya que tanto los aceites terpénicos como el CO2, se pueden recuperar y reutilizar en el mismo, consiguiendo que el impacto ambiental sea prácticamente nulo. Con el objetivo de diseñar el proceso global que permita el reciclaje de residuos de Poliestireno mediante disolución en terpenos y CO2 a alta presión, la investigación se centró en el estudio de los sistemas binarios que constituyen los diferentes pasos de éste. Inicialmente, se determinó la máxima cantidad de residuos tratados que se pueden solubilizar en los aceites terpénicos. Para ello, se estudió la influencia de la temperatura, el peso molecular y el tipo de Poliestireno (Expandido o Extruido) en la solubilidad del polímero en los terpenos. De acuerdo con esto, se comprobó que los aceites terpénicos son disolventes adecuados para llevar a cabo la disolución del Poliestireno a temperatura ambiente, independientemente del peso molecular y el tipo de procesado. Por otro lado, la solubilidad de los terpenos en CO 2 es fundamental para conseguir la separación del Poliestireno, de su disolución en los aceites terpénicos de manera eficaz. Se observó que los terpenos son totalmente solubles en CO2 a temperatura ambiente y presiones moderadas. El último sistema binario es el formado por el Poliestireno y el CO2. A pesar de que el polímero es insoluble en el gas en condiciones de presión y temperatura moderadas, la adsorción del CO2 en el Poliestireno, es responsable del hinchamiento y la plastificación del polímero. Este fenómeno conlleva la modificación de las propiedades (viscosidad, tensión interfacial y temperatura de transición vítrea) del Poliestireno, lo que afecta a su procesado en condiciones de alta presión. Una vez estudiados los sistemas binarios, es importante conocer el comportamiento de la mezcla ternaria (CO2/terpenos/Poliestireno), ya que el conocimiento aislado de los anteriores no aporta la información suficiente como para definir las condiciones de operación del proceso. Sin embargo, en base a ellos, se puede decir que la eliminación del disolvente de la mezcla es posible, ya que los terpenos son totalmente solubles en CO2 en condiciones de presión y temperatura moderadas, mientras que ocurre lo contrario en el caso del Poliestireno. Para profundizar el estudio, se determinó la influencia de la presión, la temperatura y la concentración del Poliestireno en el terpeno, en la solubilidad de los aceites seleccionados (Limoneno y p-Cimeno) en CO2. También se estudió el efecto de las variables mencionadas en la solubilidad del CO2 en la disolución o fase rica en polímero. A partir de estos resultados, se comprobó que las mejores condiciones para llevar a cabo la precipitación del Poliestireno a partir de su disolución en terpenos utilizando CO2 como antisolvente, eran aquellas en las que la densidad del antisolvente era alta (alta presión y baja temperatura) y la concentración del polímero en la disolución inicial de terpeno moderada. A continuación, se determinaron las propiedades de las mezclas a alta presión ya que para el diseño completo del proceso de reciclaje, el conocimiento del equilibrio de fases es insuficiente. Así, se observó cómo el efecto plastificante del CO 2 también afecta a las propiedades de las disoluciones, produciendo un descenso significativo de la viscosidad y la tensión superficial. Sin embargo, la temperatura de transición vítrea de las mezclas CO2/terpeno/Poliestireno que constituyen el sistema ternario es similar a la que se obtuvo en el sistema Poliestireno/CO2. Una vez conocido el sistema a partir del estudio del equilibrio y de la caracterización de las propiedades a alta presión, se llevó a cabo el proceso de reciclaje bajo dos regímenes: continuo y discontinuo. En el proceso continuo, conocido tradicionalmente como Supercritical Antisolvent (SAS), la disolución polimérica se dispersa en un recipiente que contiene CO 2. Este proceso es conocido porque permite controlar el tamaño de las partículas. Sin embargo, el Poliestireno precipitó de acuerdo a morfologías muy variadas, por lo que se estudió la hidrodinámica del sistema para intentar conocer el comportamiento de la disolución en el interior de la cámara de precipitación. El modo en el que la disolución polimérica se pone en contacto con el CO2 determina la difusión y la mezcla de fluidos y por lo tanto, la forma en la que el Poliestireno precipita. Debido a las dificultades técnicas para llevar a cabo la precipitación del polímero siguiendo una morfología preestablecida, se planteó el proceso en discontinuo, que se centra en la espumación de la disolución para producir espumas microcelulares de Poliestireno reciclado. En este caso, el CO 2 actúa no sólo como disolvente del terpeno y antisolvente del polímero, sino también, como agente espumante, lo que permite diseñar el tamaño de las celdas y su distribución, así como su densidad. La presencia del aceite terpénico favorece la transferencia de masa que está regida por la difusión del CO 2 en el polímero, pero la presencia del terpeno hace que ésta esté favorecida por la relajación de las cadenas poliméricas. El Poliestireno reciclado a través de la técnica explicada, presenta estructura microcelular, insignificantes trazas de terpeno y ningún signo de degradación. Por último, el análisis económico del proceso se llevó a cabo para determinar la viabilidad del proceso de reciclado de residuos de Poliestireno mediante la tecnología estudiada. Los costes que implica el proceso de reciclado de Poliestireno se pueden compensar si el polímero recuperado presenta mejores características que el residuo, como es el caso de la producción de espumas microcelulares. Pero también hay que destacar que la legislación reguladora de la gestión de residuos plásticos ha establecido como objetivo prioritario la eliminación de los impactos negativos que estos producen, por lo que el reciclaje debe considerarse la única alternativa. La técnica propuesta se puede llevar a cabo a temperaturas bajas, lo que disminuye los costes energéticos y el hecho de la selección selectiva del polímero evita la clasificación previa del residuo, que generalmente es muy costosa. El beneficio económico calculado, junto con las ventajas medioambientales que implica el reciclado de residuos de Poliestireno, lo convierten en un proceso muy interesante para su futura aplicación industrial. Summary Larges and increasing volumes of plastics are produced yearly due to their low price, versatility and suitability for a large number of applications and uses. Unfortunately, after its use, polymers end up in waste streams and they must be recovered in order to prevent the environmental pollution and to preserve natural sources as European waste policies are focused on. In fact, during the last decade, some positive trends are being seen concerning the recovery and recycling of polymers. Recycling of plastic wastes is attractive environmentally, but also it should present a potential value from an economic point of view in order to be a feasible process. The use of plastic wastes as source of high-added value products comes true the recycling alternative and offers a great opportunity by many reasons: the reduction of raw material, resources and energy consumption, the increase of free volume in landfills, and obviously, the environmental and social benefits. This research work is focused on the recycling of Polystyrene (PS) wastes by environmentally friendly technology for the production of high-added value products. The mentioned polymer is the target plastic since it is not biodegradable, massive amounts of energy and resource are consumed during its production, its versatility makes it commonly used but its low lifespan makes it to finish in landfills or in small pieces in the oceans. In order to promote the recycling of Polystyrene wastes, new processes should be designed to reduce the environmental impacts and costs and to improve the quality of the recovered products regarding to the traditional recycling techniques (mechanical recycling, chemical recycling or energy recovery). Thus, a technology which conforms to the Green Chemistry principle is proposed. The new recycling process of Polystyrene wastes consists on the dissolution of the polymer in natural terpene solvents to reduce its volume (decreasing transport costs), remove contaminants, dissolve specifically the polymer from the mixed plastic waste streams and avoid its degradation. Next, the separation of the terpenes from the solutions is performed by the addition of an antisolvent which makes the target polymer that precipitates while the solvent is removed. The selected antisolvent is carbon dioxide (CO2) since it is able to solubilise terpenes easily and due to its sorption in the polymer, the processing is enhanced. Furthermore, CO2 has resulted to be a very promising blowing agent which encourages its use for the production of microcellular foams, a high-added value product. Moreover, it should be highlighted that any waste stream from the process is obtained because the terpenes and the CO2 are easily recovered and reused in several cycles of the process, minimizing the environmental impact. With the aim of designing a methodology for Polystyrene recycling, the research started up by the definition of the binary systems which makes the different steps of the process. Thus, in order to determine the maximum amount of treated wastes, the influence of the temperature, the molecular weight and the source of the Polystyrene wastes (Expanded or Extruded) on its solubility in terpenes was studied. It was checked that terpenes are suitable solvents for the dissolution of the polymer at room temperature independently of its molecular weight and processing. On the other hand, the solubility of terpenes in CO2 is crucial since the success of the separation process depends on it. It was observed that terpenes are fully soluble in CO2 at room temperature and mild pressure. Finally, although Polystyrene is insoluble in CO2, the gas is absorbed into the polymer causing its swelling and plasticization. This phenomenon entails a modification of the Polystyrene properties (viscosity, interfacial tension and glass transition temperature), which affects the processing of the polymer at high pressure. Once the binary systems were defined and studied, it is important to know the behaviour of the ternary mixtures involved in the process (CO2/terpenes/Polystyrene) since the binary knowledge is not enough to select the operating conditions successfully. According to the binary system, the solvent removal is possible thanks to supercritical CO2 since it provides high solubility of terpenes and fully PS insolubility at moderated pressures and temperature. The influence of pressure, temperature and concentration in the terpenes on the solubility of the selected terpenes (Limonene and p-Cymene) in CO2 and on the sorption of CO2 into the solutions (Polystyrene/Limonene and Polystyrene/pCymene) were determined. From these results, the most suitable conditions for the precipitation of Polystyrene from its solution in terpenes using CO 2 as antisolvent were those where the antisolvent density was higher (high pressure and low temperature) and the concentration of Polystyrene in the solutions moderated. Next, the properties of the mixtures were determined since the knowledge of the phases equilibrium is too poor to define the global process. The plasticizing effect of CO2 on the polymer is also observed when it is in terpene solutions and consequently, a deep understanding of how the properties of materials and the processing parameters affect the behaviour of the mixtures is required to optimize the further development of high-quality plastic products. The study of the viscosity, interfacial tension and glass transition temperature of the mixtures was carried out. Although the viscosity and the interfacial tension of the ternary systems were significantly affected by the antisolvent, the glass transition temperature showed similar results to those obtained for the binary system Polystyrene/CO 2. Once the system was characterized and the equilibrium and properties of the ternary mixtures were determined, the recycling process was performed. Two main alternatives were proposed: continuous or discontinuous regime. In the continuous process, named Supercritical Antisolvent (SAS) the polymer solution is sprayed to a vessel previously filled with CO2. This process allows the control over the particle size. Nevertheless, Polystyrene precipitated according to a wide variety of morphologies and the knowledge of the hydrodynamic of the system was crucial to understand the process. Thus, the way in which the polymer solution is contacted with CO2 affects the diffusion and mixing of the fluids and consequently the way on the polymer precipitates. Due to the technical difficulties found along the development of the SAS process, the discontinuous one was proposed as alternative, which is focused on the foaming of the solution for the production of Polystyrene microcellular foams. In this case, CO2 behaves not only as solvent of the terpene and antisolvent of the Polystyrene, but also as blowing agent of the polymer, which allows tuning the cells size, cells size distribution and cells density of the foamed Polystyrene. The presence of the terpene promotes the mass transfer that is mainly governed by the diffusion of the CO2 into the polymeric rich phase due to the relaxation of the polymer chains. The recycled Polystyrene, obtained following this methodology, presents microcellular structure, non remarkable traces of terpene and any evidence of degradation. Finally, the economic analysis of the process was carried out to determine the feasibility of the recycling process. The traditional high costs involved in the recycling of PS wastes are balanced by the high quality recycled polymer, but also due to the mild operating conditions required to carry out the process and the selective dissolution technique which eliminates the need to sort previously the polymer from mixed plastic waste streams. The economical benefits, together with the environmental advantages show the interest of the process for industrial application. Table of contents Resumen ...................................................................................................................... 7 Summary ................................................................................................................... 11 1. 1. Plastics................................................................................................................. 5 1.1.1. Plastics production and demand........................................................................ 7 1.1.2. Plastics Wastes .................................................................................................. 9 1.1.3. Plastic Waste Management ............................................................................. 10 1.1.4. Plastic Waste Recycling ................................................................................... 13 Mechanical recycling ................................................................................................. 14 Chemical recycling .................................................................................................... 14 Energy recovery ......................................................................................................... 15 1.2. Polystyrene Production and Wastes ................................................................... 15 1.2.1. Polystyrene Recycling ...................................................................................... 17 Motivation ................................................................................................................. 27 Objectives .................................................................................................................. 28 Work plan .................................................................................................................. 29 Thesis Overview ........................................................................................................ 30 3.1. Materials............................................................................................................. 35 3.2. Experimental setups........................................................................................... 36 3.2.1. Solubility determination at atmospheric pressure .......................................... 36 3.2.2. Variable-Volume High Pressure View Cell...................................................... 36 3.2.4. High-Pressure Differential Scanning Calorimetry .......................................... 38 3.2.5. Rheometer ........................................................................................................ 39 3.2.6. Quartz Viscometer ........................................................................................... 40 3.2.3. Pendant droplet High Pressure View Cell ....................................................... 41 3.2.7. Foaming setup ................................................................................................. 41 3.2.8. Supercritical Antisolvent precipitation setup.................................................. 42 3.3. Experimental Procedures ................................................................................... 43 3.3.1. Determination of the solubility of Polystyrene in Terpene oils ....................... 43 3.3.2. High Pressure Vapour-Liquid Equilibria of Binary mixtures Terpenes/CO2 .. 44 3.3.3. High Pressure Phase Behaviour of Ternary mixtures CO2/Terpenes/Polystyrene ........................................................................................ 44 3.3.4. Glass transition temperature at high pressure ............................................... 45 3.3.5. Determination of the viscosity of Polystyrene ................................................. 46 3.3.6. Determination of the viscosity of Polystyrene solutions .................................. 46 3.3.7. Determination of high-pressure Interfacial Tension ....................................... 46 3.3.8. Foaming of Polystyrene at high pressure ........................................................ 47 3.3.9. Precipitation of Polystyrene by Supercritical Antisolvent process .................. 48 3.4. Characterisation ................................................................................................. 48 3.4.1. Thermogravimetric Analysis (TGA) ................................................................ 48 3.4.2. Thermogravimetric and Mass Spectrometer Analysis (TG-MSA) ................... 49 3.4.2. Differential Scanning Calorimetry Analysis (DSC) ......................................... 50 3.4.3. Optical Microscopy (OM) ................................................................................. 50 3.4.4. Scanning Electron Microscopy (SEM).............................................................. 51 i Table of contents 3.4.5. Determination of Molecular Weight ................................................................ 52 Graphical abstract ..................................................................................................... 57 4.1. General Background ........................................................................................... 61 4.1.1. Polystyrene/Terpene oils.................................................................................. 61 4.1.2. Terpene oils/CO2 .............................................................................................. 63 4.1.3. Polystyrene/CO2 ............................................................................................... 64 4.2. Polystyrene/Terpene oils Results........................................................................ 66 4.2.2. Determination of the solubility parameters .................................................... 66 4.2.3. Influence of the Molecular Weight on the Solubility of Polystyrene ............... 69 4.2.4. Influence of the Temperature on the Solubility of Polystyrene ....................... 71 4.2.5. Applicability to real Polystyrene wastes.......................................................... 73 4.3. Terpene oils/CO2 Results .................................................................................... 76 4.3.1. Procedure ......................................................................................................... 76 4.3.2. Literature data and correlation of binary systems Terpenoids/CO 2................ 79 4.3.3. Prediction of multicomponent mixtures and essential oils .............................. 86 4.3.4. Correlation and prediction of terpenes/CO2 using Equations of State ............ 89 4.4. Polystyrene/CO2 Results ..................................................................................... 94 4.4.1. Sorption of CO2 in Polystyrene. Plasticization................................................. 94 4.4.2. Glass transition temperature .......................................................................... 98 4.4.3. Interfacial Tension......................................................................................... 100 4.4.4. Viscosity ......................................................................................................... 102 References ............................................................................................................... 106 Graphical abstract ................................................................................................... 117 RESUMEN .............................................................................................................. 118 ABSTRACT.............................................................................................................. 119 5.1. Background of the Phases’ equilibrium ............................................................ 121 5.2. The ternary system CO2/Limonene/PS ............................................................. 122 5.2.1. Effect of pressure ........................................................................................... 125 5.2.2. Effect of temperature ..................................................................................... 126 5.2.3. Effect of concentration ................................................................................... 127 5.2.4. Sorption of CO2 into PS/Limonene solutions ................................................. 128 5.2.5. Distribution coefficient .................................................................................. 129 5.3. The ternary system CO2/p-Cymene/PS............................................................. 131 5.3.1. Ternary Diagrams ......................................................................................... 131 5.3.2. Influence of density and concentration on p-Cymene solubility in the vapour phase ....................................................................................................................... 133 5.3.3. Influence of density and concentration on CO 2 solubility in the liquid phase ................................................................................................................................. 138 5.3.4. Selection of the operating conditions ............................................................. 143 References ............................................................................................................... 147 Graphical abstract ................................................................................................... 151 RESUMEN .............................................................................................................. 152 ABSTRACT.............................................................................................................. 153 ii Table of contents 6.1. Background of viscosity and interfacial tension of the ternary mixtures CO2/Limonene/PS and CO2/p-Cymene/PS ............................................................... 155 6.2. Viscosity of mixtures CO2/Limonene/Polystyrene ............................................ 157 6.2.1. Selection of the parameters for the study of the viscosity ............................. 159 6.2.2. Concentration effect on viscosity ................................................................... 160 6.2.3. Pressure effect on viscosity ............................................................................ 162 6.2.4. Temperature effects on viscosity ................................................................... 166 6.2.5. Pressure and temperature combined effects on viscosity .............................. 169 6.2.6. Demixing determination from viscosity measurements ................................ 172 6.3. Interfacial tension of CO2/Limonene/PS and CO2/p-Cymene/PS ...................... 175 6.4. Glass transition temperature of CO2/Limonene/PS and CO2/p-Cymene/PS .... 180 Based on: ................................................................................................................. 189 Graphical abstract ................................................................................................... 189 RESUMEN .............................................................................................................. 190 ABSTRACT.............................................................................................................. 191 7.1. Supercritical Antisolvent Background ............................................................. 193 7.2. Supercritical Antisolvent Process ..................................................................... 194 7.2.1. Solubility Parameters .................................................................................... 194 7.2.2. Extraction process and recovery of solvent.................................................... 195 7.2.3. Supercritical Antisolvent Process .................................................................. 197 7.3. Foaming background ........................................................................................ 214 7.4. Feasibility of the Foaming of Polystyrene/cosolvent system with CO2 ............ 217 7.5. Statistical study of the foaming process of PS/p-Cymene solutions ................. 222 7.5.1. Characterisation ............................................................................................ 224 7.5.2. Optimisation .................................................................................................. 226 7.6. Rigorous Statistical study of the foaming process of Polystyrene/Limonene solutions .................................................................................................................. 229 7.6.1. Foaming process ............................................................................................ 230 7.6.2. Nucleation mechanisms ................................................................................. 240 7.6.3. Characterisation ............................................................................................ 244 Graphical abstract ................................................................................................... 255 RESUMEN .............................................................................................................. 256 ABSTRACT.............................................................................................................. 257 8.1. Life Cycle Assessment Background .................................................................. 259 8.2. Economical evaluation ...................................................................................... 266 iii Table of contents iv Chapter 1 INTRODUCTION Introduction 2 Chapter 1 Based on the Book Chapter: C. Gutiérrez, M.T. García, I. Gracia, A. De Lucas, J.F. Rodríguez, Polystyrene Wastes, threat or opportunity?. Environment, Energy and Climate Change I and II 3 Introduction 4 Chapter 1 1. 1. Plastics History has been frequently classified according to the materials that the man used for making his implements and other basic necessities. The most well known of these periods are the Stone Age, the Iron Age and the Bronze Age. During the last century, a new class of material has been introduced which have not only challenged the older material for their uses but has also made possible new products. Without plastics is difficult to conceive our everyday life [1]. The term plastic is derived from the Greek word plastikos meaning fit for moulding. An adequate definition of plastic is difficult to show, but generally plastic materials are synthetic of semi-synthetic organic solid widely used in the manufacture of industrial products. Attending to their molecular structure, they are mainly polymer of high molecular weight. The word polymer literally means many [2] units (mer), in which a chemical unit (monomer) is repeated itself a very large number of times to create the polymer [3]. Nowadays, the majority of commodity plastics are derived from gas or from crude oil which, after processing and refining, generate monomers that are then used in the manufacture of polymers. There are as many processing routes as kind of polymers, and several routes could be used to obtain one polymer (Figure 1.1). The huge variety of plastics makes necessary its classification in order to group by properties and applications. The most commonly used classification is based on thermoplastics (80-90 %) and thermosets (12-20%) since they constitute the majority of the market accounting for around 75% of the total polymer consumption. Thermoplastics are capable of changing shape on application of force and retaining this shape on removal of force (stress produces a nonreversible strain). They are softened when are heated above the glass transition temperature (Tg) and can be reshaped and will harden in this form upon cooling. Multiple cycles of heating and cooling can be repeated without severe damage, allowing reprocessing and recycling because they are generally linear or branched polymers which present few chemical interactions between chains. On the other hand, thermosets do not soften on heating but elevated temperatures produce chemical reactions that could harden the material into an infusible solid. Generally, they are more brittle and insoluble in organic solvents than thermoplastics because thermosets present a threedimensional structure obtained by chemical crosslinking produced after or during the processing. By this reason, thermosets can not tolerate repeated heating cycles as thermoplastics can [4]. According to their demand, the most representative thermoplastics are: Polypropylene (PP), Low Density Polyethylene (LDPE), High Density Polyethylene (HDPE), PolyVinyl Chloride (PVC), Polyethylene Terephthalate (PET) and Polystyrene (PS). The thermosets are largely represented by amino resins, polyurethanes, phenolics, polyesters, alkyd resins and epoxy resins (Figure 1.2). 5 6 Crude oil Nafta Cyclohexane Distillation Benzene or Phenol Cyclohexanone Reformer Cracking Unit CH4 CH2 CH3 Gasoline Extraction Butenes H2 C Propylene H2C CH3 Ethylene H2C Methane Styrene Ethylene oxide Ethylene glycol Butadiene CH2 Polyglycols Polyglycols p-Xylene; m-Xylene Ethyl Benzene Cumene Phenol Styrene Isopropyl alcohol Acetone Propylene oxide Propylene glycol Caprolactam Phthalic anhydride CH Ethylene dichloride o-Xylene Mixed Xylenes Toluene Benzene Benzene HC Acetylene Alkyl resins Polyesteres Polyamides Synthetic rubbers Polymethyl methacrylate Polypropylene Polystyrene Polyester fibres and films Polyurethanes Polyethylene Polyvinyl chloride Introduction Figure 1.1. Conventional production routes for polymers raw materials to commodity plastics. Chapter 1 9000 8000 Production (kTonnes) 7000 6000 5000 4000 3000 2000 1000 0 P LD P E LD DP /M E/L PE D H PP PS C PV T PE AN S/S AB MA PM PC R PU s her Ot Type of plastics Figure 1.2. Demand of most representative thermoplastics in Europe 2012. 1.1.1. Plastics production and demand In the second half of the 20th century, plastics became one of the most-universallyused and multipurpose materials in the global economy due to their low price and versatility. Today, plastics are used in more and more applications and they have become essential to our modern economy [5]. Among the most important properties are: versatility, low price, lightweight, durability and strength. Due to their excellent characteristics, the plastics industry has benefited from 50 years of growth with a year on year expansion of 8.7% from 1950 to 2012. The production and consumption of plastic materials have been growing steadily. For instance, the production of plastics rose to 288 million tonnes from 2011 to 2012 which means an increase of 2.8 %. Nevertheless in Europe and according to the general economic situation, plastics production decreased by 3% from 2011 to 2012 (Figure 1.3) [6]. 7 Introduction World Plastic Production Latin America 4.9% NAFTA 19.9% 300 World Plastic Production European Plastic Production Biopolymers Plastic Production (Mtonne) 250 200 Concers for wildlife 100 Accumulation in natural habitats France 9.37% in 95 60 rt s ) o e ( f en 19 t an 70 s ) g l em 19 co po Germany Spain UK Italy France Benelux Other EU Benelux 4.48% 1990 2000 2010 Other EU 34.42% UK 7.54% Spain 7.13% Germany 23.63% in by wi ld re rs t Fi lif s( i rd ab eb se of sd ts 1980 Year ed (1 rd ed er re ov ri s isc ed tic gu yl en as Pl 1970 en EU ts P t ro ack ag du in ce g d b y d ir M ect i un v i ci e l e Pl g pa as tic l S isla t ol de id io n br is W re as po t es rt e (1 d 99 in 0s de ) ep se a( 20 00 ) 1960 4) 1950 op China 23.9% Italy 13.44% 0 pr Japan 4.9% Europe Plastic Production Product user diversification ly Middle East, Africa 7.2% China Rest of Asia NAFTA Latin America Europe CIS Middle East, Africa Japan 50 Po CIS 3% Rest of Asia 15.8% Concerns for human life 150 Europe 20.4% Figure 1.3. Plastics production from 1950 to 2012. The production includes thermoplastics, polyurethanes, thermosets, elastomers, adhesives, coating, sealants and fibres. But PET, PA and polyacryl-fibres are not included [6]. The first plastic producer country around the world is China, followed by Europe and the United States of America, Canada and Mexico (NAFTA). European countries produced 50 Mtonnes in 2012 and the main producers were Germany, Italy and France. UK and Spain got the fourth and fifth position, respectively. In Europe, packaging purposes are the largest application sector for the plastic industry and represent 39.4% of the total plastics demand. In the second position, building and construction sector reaches 20.3% followed by the automotive sector, electrical and electronic applications. Other application sectors such as appliances, household and consumer products, furniture and medical products comprise a total of 22.4% of the European plastics demand (Figure 1.4). 8 Chapter 1 Automotive Electrical&Electronics 3.76 Mtonnes 2.52 Mtonnes Agriculture Building & Construction 1.93 Mtonnes 9.32 Mtonnes Others 10.28 Mtonnes Packaging Building & Construction Automotive Electrical & Electronics Agriculture Others Packaging 18.08 Mtonnes Figure 1.4. Main sectors of application of the plastics industry in Europe 2012. 1.1.2. Plastics Wastes It is obvious that plastics are used in our daily lives in a great number of applications, but at the end of their useful life, polymers enter in waste streams as either post-consumer waste or industrial scrap. Households, distribution and industry sectors are the main sources of the wastes. But also, the automotive sector, the electrical and electronic, as well as building and construction provides important amounts of plastic wastes (Figure 1.5). 6000 PE PP PVC ABS PS PU PC Others Total Plastic Waste (kTonnes) 5000 4000 3000 2000 1000 0 Municipal solid wastes Automotive Agriculture Construction Electrical&Electronic Origin and Type of Post-consumer Plastic Wastes Figure 1.5. Types of plastic originated depending on the origin of wastes. 9 Introduction Plastic wastes have been involved in the municipal solid waste considerably, concentrated in packaging, films, covers, bags, containers and so on with widely used in our daily life. According to Figure 1.5, municipal wastes cover the largest volume of plastics wastes. Packaging materials make up the largest contribution to polymer waste in the stream and they are mainly constituted by LDPE/LLDPE, PP, HDPE, EPS/PS, PVC and PET. The polyolefins used for packaging materials are often discarded after a single use, which results in a large amount of polymer wastes [7]. In the automotive sectors, plastics have increasingly been used to replace metals for many components because of their greater processability, lighter weight and corrosion resistance. The percentage by weight of polymers in the average European car has risen in the last years, in contrast with packaging applications. Polymers found in automotive wastes are PP, ABS, HDPE, PP and PU, PVC. The plastic wastes coming from the construction are mainly PVC and thermosetting resins together with acrylics or PP. The nature of the plastics generated in the large industry is mainly PE from sheet used exclusively for all secondary packaging. LDPE is the predominant plastic from the agricultural plastic wastes. And eventually, electrical and electronic wastes are a significant contributor of ABS poly(acrylonitrile-co-butadiene-co-styrene), polycarbonate and modified PP. From the comparison of Figure 1.2 and 1.4, significant differences between the plastic demand and plastic waste generation can be observed. The most demanded plastic in Europe in 2012 was Polypropylene, while the Polyethylene was the most common plastic presents in waste streams. Although the vast majority of the industrial wastes are recycled with the processing, the rests are usually sent for reprocessing. On the other hand, the majority of post-consumer plastics waste reaches the environment and hence the emphasis in polymer waste management is on this type of waste stream. 1.1.3. Plastic Waste Management European waste policy has regulated the management of the municipal solid wastes through “environmental actions plans and a framework of Legislation” during the last 30 years (European Commission 2010). The target of the environmental policy is the reduction of waste negative impacts. In 2005, a long-term strategy on waste management which started a new era of EU waste policy. The new strategy stated from the EU’s Sixth Environment Action Programme has been reflected in the Waste Framework Directive (WFD-2008/98/EC). The new Directive (WFD2008/98/EC) is focused on the promotion of a recycling society with the targets for “EU Member States to recycle 50% of their municipal wastes and 70% of construction wastes by 2020” [8]. Wastes management legislation in Europe differs from country to country; however, basic principles and restrictions are in accordance 10 Chapter 1 with the European Landfill Directive and other important EU legislation in the field of wastes management. In this regard, since the last years a positive trend is growing in the recovery and recycling of plastics in Europe. In 2012, post-consumer plastic wastes reached 25.2 Mtonnes, of which 62% of plastics were recovered and only 38% were ended up in landfills. It is important to highlight that almost 70% of plastic packaging wastes were recovered, 34.5% went through energy recovery (incineration or additive as refuse derived fuel), 34.2% was mechanically recycled and only 0.7% went to feedstock recycling [6]. It should be outlined that the recycling of plastic wastes is attractive not only environmentally, but also from an economic point of view due to their high potential value [9]. European Commission marked a shift away from thinking about wastes as an unwanted burden to seeing it as a valued resource (European Commission 2010). But the recycling will be only economically advantageous if the design of the process consider the amount of energy necessary to produce the virgin polymer plus the energy to dispose equal to the energy to recover the plastic wastes plus the energy during the reprocessing. The high cost would be balanced by the reduction cost of landfilling and of course by the social benefit [10] since the recycling of plastics greatly contributes to the preservation of the natural resources. According to the data, recycling methods are being established as priority techniques for the recovery of polymers from plastic wastes. But the recycling of materials requires the separation of the different polymers in order to recover plastics of acceptable quality. Plastic recycling streams are made of dirty mixed polymers including dyes, fillers, flame retardant and other additives [9]. The use of polymers in different market sectors complicates the identification and separation of waste streams because their use in one sector does not guarantee that they will appear as waste in the same sector [4]. By this reasons, the separation of different polymers by type is crucial. But also, the differences among the melting temperature and some chemical incompatibilities hinder the mixture of plastics previously to its classification, especially if the target recycled product should present high value [11]. Adequate and cost-effective techniques are demanded for the identification of plastics since the recycling processes are priority as it was shown in Figure 1.6 [12]. Separation is further complicated by the current design of products in which polymers are commingled with other materials and polymers are currently sorted and separated manually or mechanically which means its high cost and inefficiency [13]. Considerable work has been undertaken to develop automated technologies that can separate mixed waste streams according to polymer type or to remove foreign matter (contaminants) from the waste stream. Nevertheless, plastic exhibit similar properties which make the automatic classification a challenge. 11 12 Washing Plasticisers Stabilisers Thermosets Plastic wastes Flame retardant Pigments Thermoplastics Classification Infrared Raman Thermal Analysis Fluidization Spectroscopy Laser Manual SELECTIVE DISSOLUTION Energy Recovery Chemical Recycling Mechanical Recycling Precipitation Co-Combustion Incineration Chemolysis Pyrolysis Hydrogenation Gasification Thermoplastics Thermosets Hydrolysis Glycolysis Methanolysis 400-900 ºC Fluidised bed, rotatory kiln, tubular crackers Bubble column 20 MPa, 500 ºC Fluidised-bed, fixed-bed, pneumatic transport reactor 15-30 MPa, 800-600 ºC Melting Extrusion Grinding Milling RECYCLED POLYMER MONOMERS WAX, OIL, GAS, ENERGY SYNCRUDE SYNTHESIS GAS RECYCLED POLYMER ADDITIVES Introduction Figure 1.6. Routes for the management of the plastic wastes: from the sorting to the recycling. Chapter 1 Figure 1.6 shows some of the developed techniques for the identification and sorting of different types of plastics; the substantial development occurred in the last decade has achieved a so good classification of plastic wastes that only contain small amounts of other types of polymers [14]. Although, gravity separation method based on air or centrifugal force was proposed by Choi et al. [15], the most employed techniques are related with spectroscopy: Laser [16], Raman [17], Infrared [17] or fluorescence [9]. FTRaman has been proved as the most rapid and selective technique to recognize the most usual polymers. But generally, the combination of different analytical techniques is required to sort perfectly the plastic wastes streams. This fact makes these technologies very expensive and its efficiency is always limited by the feed characteristics. By these reasons, the choice of the most appropriate will depend on the complexity of the polymer mixture, its physical form, the quality or the additives and/or contaminants [4]. The proposal of a general methodology for the classification of plastic wastes is still nowadays a challenge. There is no clear relationship between quantities of plastics produced and quantities for plastic wastes, because of the lag between production and disposal depends mainly on the lifespan of each product. Post-consumer plastics wastes may arise from a host of products or applications each with differing life cycles and therefore they should be treated individually. While containers and packaging wastes generally have lifespans under a year; durable and non-durable plastics can reach lifespans estimated at five years for transportation applications; 10 years for furniture, housewares, electric and electronic; and 50 years or more for building and construction materials [18]. Generally and according to this data, packaging dominates the wastes generated from plastics, covering 62.2 % of the total. Other applications like building and construction, electrical and electronic products and agriculture count for 5 till 6% each. It should be more appropriate to divide in new categories the waste streams in which plastics may be found (Figure 1.5). Years of research, study and testing have resulted in a numbered of treatment, recycling and recovery methods that could be feasible and economically advantageous. Important efforts are making to manage efficiently the plastic products at the end of their service life, new routes are improving constantly and fewer of them are ending up in landfills. 1.1.4. Plastic Waste Recycling According to Figure 1.6, plastic wastes management can be performed mainly by three different ways: mechanical recycling, chemical recycling and energy recovery [19]. A fourth approach was defined by Scott [20] as the biological recycling but due to the high costs, complex procedures and environmental damages involve during its development, this alternative is not generally considered [10]. Each method provides 13 Introduction a unique set of advantages that make it particularly beneficial for specific applications [21]. Next, a brief summary of the different alternatives is shown. Mechanical recycling Separation of different polymers is particularly important for mechanical recycling because processing mixed materials would produce recycled products of low quality, which could only be used in a limited number of applications. Hence, mechanical recycling is really best suited to clean plastic wastes such as packaging materials. Mechanical recycling of mixed wastes necessarily starts with the manual sorting process because it can only be performed on single-polymer plastics and the mixed or contaminated wastes can not be recycled following this method. This fact implies high costs and low quality products when the separation is not as efficient as is required. Depending on the type of polymer, the mechanical recycling follows different processes. Thermosets are grinded and particulated for its re-used as filler in new materials since they can not be re-melted. They are generally used as additives to improve the properties of other polymers providing better values of modulus, elongation at break or impact strength. On the other hand, thermoplastics are remelted, extruded and pelletized to be sold as a raw material for further processing. Although thermoplastics are soften above their glass transition temperature, degradation and heterogeneity of wastes should be considered. Polymerization reactions are irreversible but the heat provided during the mechanical recycling could cause photo-oxidation. Consequently, length or branching of polymers could occur and the polymer would present worse properties as a result of the ageing. Despite this, mechanical recycling has been the oldest practice by common plastics manufacturers [22]. Chemical recycling Chemical recycling, is based on the conversion of polymers into monomers (monomer recycling) or useful petrochemicals (feed-stock recycling). Generally, polyolefin plastics with a massive production have been subjected to recycled chemically to produce a mixture of hydrocarbons that could be useful as fuel [21] [23]. Several processes are included in the chemical recycling and the most relevant are explained below (Figure 1.6). Gasification is the partial oxidation of hydrocarbons in the presence of low oxygen levels. The process is typically carried out at temperatures from 800 ºC up to 1600 ºC, pressures of 15-30 MPa and in presence of a gasifying agent (air, oxygen, steam, fuel gas). The main products of gasification are syngas [24]. 14 Chapter 1 Once plastic wastes are depolymerised, they can be hydrogenated to produce bitumen and a synthetic crude oil, known as syncrude which present a very high added value. The syncrude is refined for its use in the petrochemical industry. Pyrolysis consists on the thermal decomposition of the plastic wastes at temperature from 350-700 ºC in the absence of oxygen or other gasifying gases. The polymers decompose to their monomers, oligomers and other organic substances that can be collected separately and used as a feedstock or for energy generation [25]. Finally, chemolysis is a treatment which uses solvolytic processes (hydrolysis, glycolysis or alcoholysis) to recycle and convert plastics into their basic monomers for repolymerisation into new plastics. This method allows up to 90% of the plastic (by weight) to be re-used but it works best with homogeneous plastic types. Energy recovery When the recycling of plastic wastes is not feasible, they could be used in energy recovery due to its high calorific values. The high calorific value of plastics underscores the need for an alternative energy recovery technology that is affordable, efficient, safe and user-friendly. The two ways to obtain energy from plastic wastes are the incineration and the co-combustion. Incineration without energy recovery would be also possible but is not acceptable from the sustainability point of view because it only reduces the amount of wastes, but waste valuable nonrenewable resources and it can not be defined as a recycling option. Furthermore, the incineration of polymer provokes important social opposition [4]. Despite plastics recycling is increasing and many different alternatives are proposed, important efforts are still required to recycle efficiently the plastic wastes. Mechanical recycling is only economically feasible if the plastic wastes are classified in a single polymer stream and energy recovery must also separate PVC from the rest of the polymers due to the formation of dioxins. The fact that most of the recycling options require a severe previous classification of polymers by type is a challenge that nowadays constrains the recycling. Although the recycling has been established as a priority option, the mentioned high costs related with the sorting process limits its application. 1.2. Polystyrene Production and Wastes In this work, we are focused on the recycling of Polystyrene wastes because in the last years it has been considered a plague on the environment taking more space in landfills than paper. The first commercial production of Polystyrene began in the early 1930s by the Farben Company in Germany. Polystyrene is the polymerization product of styrene monomer, which is a colourless liquid with a strong odour derived from petroleum and natural gas by-products (Figure 1.1) [26]. 15 Introduction n Styrene Polystyrene Figure 1.7. Styrene polymerization Generally, Polystyrene is a clear, hard, glassy material with a bulk density between 0.94-1.05 g/cm3. It founds application in packaging, food and beverage containers, building and insulation. It is widely used since it is very cheap, light, and versatile also it presents low dielectric constant and low thermal conductivity. Unfortunately, due to its resistance to biodegradation, PS plastic wastes have become one of the major problems to the environment. PS is made from petroleum, a non-renewable resource, through a very complicated process. From oil to monomer, monomer to polymer, and polymer to final product (Figure 1.7), not only are massive amounts of energy and resource consumed, but also serious environmental impact is generated by the productive processes [27] [28]. The Environmental Protection Agency (EPA) established PS as the fifth largest source of hazardous wastes. According to Kyoto protocol, PS foams were banned to the ozone-depleting CFC gases used as blowing agents. Nowadays, several governments are considering the extension of the ban on PS foam containers because its impact on marine pollution. Scientists are concerned that plastics debris in the ocean can transport toxic substances which may end up in the food chain causing potential harm to ecosystem and human health (UNEP 2011). Regarding the cited restrictions, the traditional disposal of plastic wastes in landfills is not an option due to the high costs, legislative pressures and public concern on resource conservation. PS should be considered a raw material with high potential value and its recycling a priority, not only for the environmental issues but also from an economical point of view. Nevertheless, its excellent properties entail some environmental inconveniences. It is extremely lightweight and due to the high volume/weight ratio, it does not seem economically viable to store or transport nor is an attractive material for collection taking up huge volume in landfills. But if PS is landfilled there is an important loss of the potentially valuable materials. Its versatility means a wide variety of applications which makes often the PS wastes a mixture of different grades and types of polymer. This fact could result in variable and low quality recycled products, which limits the further uses of the recovered material. On the other hand, PS wastes (especially EPS) are frequently contaminated by food and they are generally not recycled because clean wastes are required in order to recover high quality plastics. It should be emphasise that recycled plastics should 16 Chapter 1 not present a lower grade than the "virgin" polymer because it could prevent its use in some applications, such as that for food packaging. Eventually, it should be underlined that PS generally is a one-time use material and its cost per unit is around cents. The world’s oversupply of this virgin polymer, due to the excess of production capacity, has driven its price down. So it is a challenge the recycling of PS wastes to achieve a high market value product from an easy and cheap process since the cost for collecting, sorting, cleaning and reprocessing often become excessive [29]. The main idea to overcome the economical boundaries is the recovery of the material with enhanced properties and at same time with new uses [10]. Mechanical recycling has been selected as the most adequate way to recover styrenic plastic wastes from packaging applications. But during the re-melting and compression of PS, their chemical structure, long-term stability or mechanical properties could be altered [30]. Although the recycling of the cited polymer has been highly promoted, nowadays the rate of PS recycling is very low in comparison to the recycling rate of all other plastic wastes. 1.2.1. Polystyrene Recycling According to the different aspects about the recovery of PS wastes, all the scientific efforts should be focused on the recycling since it allows saving energy and raw materials. This fact is crucial due to the existing situation that Europe relies upon imports of scarce resources (European Commission, 2010). New processing schemes should be explored in order to reduce the cost of the traditional and expensive recycling processes mentioned before. Dissolution and shrinking has been proposed as one of the cheapest and most efficient methods for the recycling of PS. The selection of the most suitable solvent is the crucial step for the selective recycling of polymers by dissolution [31]. The main advantages of the dissolution of polymeric wastes are shown below: the dissolution of PS in a suitable solvent will cause considerable volume reduction (it could be achieved it a volume reduction of more than 100 times). any insoluble contaminants will be dissolved and they could be removed by filtration, obtaining a cleaner polymer that could be reprocessed. the selection dissolution allows the separation of plastics from other types of waste and polymers non-soluble depending on their chemical nature. For instance, a single solvent can be used to dissolve several unsorted polymers at different temperatures or even it can dissolve selectively only one polymer. the polymer does not suffer further degradation, unless heat is used for the dissolution of the plastic wastes. 17 Introduction In view of these numerous advantages, the design of a process for the recycling of PS wastes by dissolution can be performed. The dissolution of PS wastes has been studied but focused on the thermal recycling (pyrolysis) [32]. With this aim, it was proposed the use of several solvents: biodiesel [33], solvents derived from the pyrolysis of the PS wastes [34], aliphatic, cyclic, and aromatic solvents [35] or even automobile lubricating oil wastes [36]. In all cases the addition of solvent provided solutions with a high specific heat. The heat of combustion of PS is approximately 41 MJ/kg, comparable to the traditional value for oil. In a proper designed combustion device complete combustion could be achieved resulting in water, carbon dioxide and trace levels of ash. The use of the aforementioned solvents has been focused on the dissolution of PS wastes in low cost solvents but with high specific heat in order to be pyrolysed. Nevertheless, after the pyrolytic treatment only the energy is recovered but not the polymer. However, it can be checked that the concept of dissolution could be useful to propose new methodologies for the recycling of PS wastes. Polymer dissolution plays a key role in many industrial applications among them, recycling of polymers by dissolution is becoming a green alternative to perform the process. A single solvent can be used to dissolve several unsorted polymers at different temperatures or even it can dissolve selectively only one [37-41]. But the dissolution of polymeric materials is different from non-polymeric since in polymers, the dissolution process is generally controlled by the external mass transfer resistance through a liquid layer and they are not dissolved immediately [42]. Generally, polymer dissolution processes show two steps: the disentanglement of polymer chains when the solvent penetrates and the polymer swelling. Thus, the bulk transport process implies the movement of disentangled chains from the polymer to solvent rich phase [43]. Figure 1.8 shows a scheme about the structure of the surface layers of glassy polymers during their dissolution in a pure solvent. Polymer Infiltration Layer Solid Swollen Layer Gel Layer Liquid Layer Solvent Figure 1.8. Schematic diagram of the composition of the surface layer. As can be observed in Figure 1.8, a polymer contains free volume (channels and holes of molecular dimensions) where solvent goes in firstly. From those empty spaces the diffusion of the solvent starts and creates a solid swollen layer. Next, the gel layer is ensued from the swollen polymer, which contains material in a rubberlike state and a liquid layer surrounding. 18 Chapter 1 Next, solvent removal produces amorphous or crystalline porous materials depending on the experimental conditions, such as polymer concentration, molecular mass or temperature [44]. The simplest way to eliminate and recovery the solvent is evaporation or distillation (vacuum or steam) [45, 46], but high temperature can promote the degradation of the polymer chains [47]. As an alternative, the addition of a proper antisolvent makes the polymer precipitate while the solvent is solubilised in the antisolvent and both can be recovered independently [37, 48]. The ratio solvent/antisolvent, the concentration of the polymer in the initial solution, the final solvent content in the recovered polymer or the form of the precipitated are some of the variables to consider during the design of the recycling of polymers by dissolution and antisolvent process. Recycling processes only are valuable if they are also economically beneficial, it means, the recycled material, prepared from post-consumer waste plastic, should present high-quality. In general, polymers have no competing materials in terms of weight, ease of processing, economy and versatility [6]. Particularly, microcellular foams have enjoyed a continuously growing market due to their high impact strength, toughness and thermal stability, as well as low dielectric constant and thermal conductivity; therefore they are a high-quality product. Microcellular foams are defined as foams characterized by cell sizes of less than 10 microns and cell densities greater than 109 cells/cm3 [49]. Plastic foams have been produced by a variety of processes such as batch foaming, extrusion foaming, and injection foam molding. Table 1.1. summarizes the general technologies used to make polymer foams and the target applications as a function of their . Table 1.1. Polymeric Foams methodology. Technology Reactive foaming Extrusion Injection Mold Polymers Applications Thermosets: Polyurethanes and phenolics Thermoplastics: Polystyrene, Polyvinyl Chloride, Polyethylene, Polypropyelene Thermoplastics: Polystyrene, Polypropylene, Polyethylene Construction, automotive, furniture, packaging Food, construction, packaging, medical Automotive In general, thermosets are foamed by means of reactive methodology and thermoplastics by extrusion or injection [50]. Table 1.1 shows that the cellular structure in plastics may be produced mechanically, chemically or physically [51]. Thus, blowing agents are generally classified in chemical or physical blowing agents. Polystyrene is a thermoplastic generally foamed by physical methodologies using physical blowing agents. 19 Introduction An ideal physical blowing agent should be environmentally acceptable, nonflammable, adequately soluble, stable in the process, low toxicity, low volatility, low vapour thermal conductivity, low molecular weight and low cost [52]. Supercritical carbon dioxide (scCO2) has resulted to be a very promising blowing agent for the replacement of volatile organic compounds especially used in the production of foams from glassy polymers [53-56]. The main advantage of CO2 is that it is relatively inert, non flammable, cheap and presents tuneable physicochemical properties near its critical point (Tc: 31.1 ºC, Pc: 73.8 bar) [57]. Furthermore, CO2 can be used as antisolvent for the polymer and solubilise the solvent achieving the precipitation of the Polystyrene from its solution on microcellular foams like [58]. 20 Chapter 1 References [1] J.A. Brydson, Plastics Materials, Elsevier Science, 1999. [2] P.P. Council, Polystyrene Packaging Council, in: A. Spangenberg (Ed.), Polystyrene Packaging Council, n.d. [3] P. Ghosh, Polymer Science and Technology: Plastics, Rubbers, Blends and Composites, Tata McGraw-Hill, 2001. [4] A. Azapagic, A. Emsley, I. Hamerton, Polymers: The Environment and Sustainable Development, Wiley, Guildford, UK, 2003. [5] D.W. van Krevelen, K. te Nijenhuis, Properties of Polymers: Their Correlation with Chemical Structure; their Numerical Estimation and Prediction from Additive Group Contributions, Elsevier Science, 2009. [6] PlasticsEurope, PlasticsEurope, in: T.P. Portal (Ed.). [7] C. Borsoi, L.C. Scienza, A.J. Zattera, Characterization of composites based on recycled expanded polystyrene reinforced with curaua fibers, Journal of Applied Polymer Science, 128 (2013) 653-659. [8] T. Zelenović Vasiljević, Z. Srdjević, R. Bajčetić, M. Vojinović Miloradov, GIS and the analytic hierarchy process for regional landfill site selection in transitional countries: A case study from Serbia, Environmental Management, 49 (2012) 445458. [9] S. Brunner, P. Fomin, D. Zhelondz, C. Kargel, Investigation of algorithms for the reliable classification of fluorescently labeled plastics, in, Graz, 2012, pp. 1659-1664. [10] A.F. Ávila, M.V. Duarte, A mechanical analysis on recycled PET/HDPE composites, Polymer Degradation and Stability, 80 (2003) 373-382. [11] T. Carvalho, F. Durão, C. Ferreira, Separation of packaging plastics by froth flotation in a continuous pilot plant, Waste Management, 30 (2010) 2209-2215. [12] K. Inada, R. Matsuda, C. Fujiwara, M. Nomura, T. Tamon, I. Nishihara, T. Takao, T. Fujita, Identification of plastics by infrared absorption using InGaAsP laser diode, Resources, Conservation and Recycling, 33 (2001) 131-146. [13] S.S. Martínez, J.M.L. Paniza, M.C. Ramírez, J.G. Ortega, J.G. García, A sensor fusion-based classification system for thermoplastic recycling, in, Loughborough, Leicestershire, 2012, pp. 290-295. [14] B. Luijsterburg, H. Goossens, Assessment of plastic packaging waste: Material origin, methods, properties, Resources, Conservation and Recycling, (2013). [15] W.Z. Choi, J.M. Yoo, E.K. Park, Separation of individual plastics from mixtures by gravity separation processes, in, San Diego, CA, 2006, pp. 459-468. [16] V.K. Unnikrishnan, K.S. Choudhari, S.D. Kulkarni, R. Nayak, V.B. Kartha, C. Santhosh, Analytical predictive capabilities of Laser Induced Breakdown Spectroscopy (LIBS) with Principal Component Analysis (PCA) for plastic classification, RSC Advances, 3 (2013) 25872-25880. [17] J. Anzano, M.E. Casanova, M.S. Bermúdez, R.J. Lasheras, Rapid characterization of plastics using laser-induced plasma spectroscopy (LIPS), Polymer Testing, 25 (2006) 623-627. [18] H. Alter, The recovery of plastics from waste with reference to froth flotation, Resources, Conservation and Recycling, 43 (2005) 119-132. [19] K. Hamad, M. Kaseem, F. Deri, Recycling of waste from polymer materials: An overview of the recent works, Polymer Degradation and Stability, 98 (2013) 28012812. [20] G. Scott, "Green" polymers, Polymer Degradation and Stability, 68 (2000) 1-7. 21 Introduction [21] S.M. Al-Salem, P. Lettieri, J. Baeyens, The valorization of plastic solid waste (PSW) by primary to quaternary routes: From re-use to energy and chemicals, Progress in Energy and Combustion Science, 36 (2010) 103-129. [22] M. Al Shrah, I. Janajreh, Mechanical recycling of cross-link polyethylene: Assessment of static and viscoplastic properties, in, Ouarzazate, 2013, pp. 456-460. [23] G. De La Puente, U. Sedran, Recycling polystyrene into fuels by means of FCC: Performance of various acidic catalysts, Applied Catalysis B: Environmental, 19 (1998) 305-311. [24] V. Wilk, H. Hofbauer, Conversion of mixed plastic wastes in a dual fluidized bed steam gasifier, Fuel, 107 (2013) 787-799. [25] W. Kaminsky, M. Predel, A. Sadiki, Feedstock recycling of polymers by pyrolysis in a fluidised bed, Polymer Degradation and Stability, 85 (2004) 1045-1050. [26] W.C. Teach, G.C. Kiessling, Polystyrene, Reinhold Publishing Corporation, 1960. [27] A.L. Andrady, Plastics and the Environment, Wiley, 2003. [28] J. Brandrup, Recycling and recovery of plastics, Hanser Publishers, 1996. [29] C.A. Ambrose, R. Hooper, A.K. Potter, M.M. Singh, Diversion from landfill: Quality products from valuable plastics, Resources, Conservation and Recycling, 36 (2002) 309-318. [30] F. Vilaplana, A. Ribes-Greus, S. Karlsson, Degradation of recycled high-impact polystyrene. Simulation by reprocessing and thermo-oxidation, Polymer Degradation and Stability, 91 (2006) 2163-2170. [31] C. Gutiérrez, M.T. García, I. Gracia, A. De Lucas, J.F. Rodríguez, A practical approximation to design a process for polymers recycling by dissolution, Una aproximación práctica para el diseño de un proceso de reciclado de polímeros mediante disolución, 68 (2011) 181-188. [32] J.M. Arandes, J. Ereña, M.J. Azkoiti, M. Olazar, J. Bilbao, Thermal recycling of polystyrene and polystyrene-butadiene dissolved in a light cycle oil, Journal of Analytical and Applied Pyrolysis, 70 (2003) 747-760. [33] Y. Zhang, S.K. Mallapragada, B. Narasimhan, Dissolution of waste plastics in biodiesel, Polymer Engineering and Science, 50 (2010) 863-870. [34] Y. Kodera, Y. Ishihara, T. Kuroki, S. Ozaki, Feedstock Recycling of Plastics, in: M. Müller-Hagedorn, H. Bockhorn (Eds.) Solvo-Cycle Process: AIST's New recycling process for used plastic foam using plastics-derived solvent, Karlshrue, 2005, pp. 217-222. [35] A. Karaduman, E.H. Imşek, B. Çiçek, A.Y. Bilgesü, Thermal degradation of polystyrene wastes in various solvents, Journal of Analytical and Applied Pyrolysis, 62 (2002) 273-280. [36] S.S. Kim, J. Kim, J.K. Jeon, Y.K. Park, C.J. Park, Non-isothermal pyrolysis of the mixtures of waste automobile lubricating oil and polystyrene in a stirred batch reactor, Renewable Energy, 54 (2013) 241-247. [37] G. Pappa, C. Boukouvalas, C. Giannaris, N. Ntaras, V. Zografos, K. Magoulas, A. Lygeros, D. Tassios, The selective dissolution/precipitation technique for polymer recycling: A pilot unit application, Resources, Conservation and Recycling, 34 (2001) 33-44. [38] D.S. Achilias, A. Giannoulis, G.Z. Papageorgiou, Recycling of polymers from plastic packaging materials using the dissolution-reprecipitation technique, Polymer Bulletin, 63 (2009) 449-465. 22 Chapter 1 [39] A.J. Hadi, G.F. Najmuldeen, K.B. Yusoh, Dissolution/reprecipitation technique for waste polyolefin recycling using new pure and blend organic solvents, Journal of Polymer Engineering, 33 (2013) 471-481. [40] E.M. Kampouris, D.C. Diakoulaki, C.D. Papaspyrides, SOLVENT RECYCLING OF RIGID POLY(VINYL CHLORIDE) BOTTLES, Journal of Vinyl and Additive Technology, 8 (1986) 79-82. [41] E.M. Kampouris, C.D. Papspyrides, C.N. Lekakou, A model recovery process for scrap polystyrene foam by means of solvent systems, Conservation and Recycling, 10 (1987) 315-319. [42] B.A. Miller-Chou, J.L. Koenig, A review of polymer dissolution, Progress in Polymer Science (Oxford), 28 (2003) 1223-1270. [43] B.S. Kong, Y.S. Kwon, D. Kim, Theoretical and experimental analysis of polymer molecular weight and temperature effects on the dissolution process of polystyrene in ethylbenzene, Polymer Journal, 29 (1997) 722-732. [44] J. De Rudder, H. Berghmans, J. Arnauts, Phase behaviour and structure formation in the system syndiotactic polystyrene/cyclohexanol, Polymer, 40 (1999) 5919-5928. [45] K. Hattori, S. Shikata, R. Maekawa, M. Aoyama, Dissolution of polystyrene into p-cymene and related substances in tree leaf oils, Journal of Wood Science, 56 (2010) 169-171. [46] S. Shikata, T. Watanabe, K. Hattori, M. Aoyama, T. Miyakoshi, Dissolution of polystyrene into cyclic monoterpenes present in tree essential oils, Journal of Material Cycles and Waste Management, 13 (2011) 127-130. [47] M.T. García, G. Duque, I. Gracia, A. De Lucas, J.F. Rodríguez, Recycling extruded polystyrene by dissolution with suitable solvents, Journal of Material Cycles and Waste Management, 11 (2009) 2-5. [48] J.G. Poulakis, C.D. Papaspyrides, Recycling of polypropylene by the dissolution/reprecipitation technique: I. A model study, Resources, Conservation and Recycling, 20 (1997) 31-41. [49] J.B. Bao, T. Liu, L. Zhao, G.H. Hu, X. Miao, X. Li, Oriented foaming of polystyrene with supercritical carbon dioxide for toughening, Polymer (United Kingdom), (2012). [50] S.T. Lee, N.S. Ramesh, Polymeric Foams: Mechanisms and Materials, Taylor & Francis, 2004. [51] R. Gendron, Thermoplastic Foam Processing: Principles and Development, Taylor & Francis, 2004. [52] D. Klempner, V. Sendijareviʹc, R.M. Aseeva, Handbook of Polymeric Foams and Foam Technology, Hanser Publishers, 2004. [53] S.G. Kazarian, Polymer processing with supercritical fluids, Polymer Science Series C, 42 (2000) 78-101. [54] A.I. Cooper, Porous materials and supercritical fluids, Advanced Materials, 15 (2003) 1049-1059. [55] D.L. Tomasko, A. Burley, L. Feng, S.K. Yeh, K. Miyazono, S. Nirmal-Kumar, I. Kusaka, K. Koelling, Development of CO2 for polymer foam applications, Journal of Supercritical Fluids, 47 (2009) 493-499. [56] D.L. Tomasko, H. Li, D. Liu, X. Han, M.J. Wingert, L.J. Lee, K.W. Koelling, A Review of CO2 Applications in the Processing of Polymers, Industrial and Engineering Chemistry Research, 42 (2003) 6431-6456. [57] K. Wu, J. Li, Precipitation of a biodegradable polymer using compressed carbon dioxide as antisolvent, Journal of Supercritical Fluids, 46 (2008) 211-216. 23 Introduction [58] D.J. Dixon, K.P. Johnston, Formation of microporous polymer fibers and oriented fibrils by precipitation with a compressed fluid antisolvent, Journal of Applied Polymer Science, 50 (1993) 1929-1942. 24 Chapter 2 MOTIVATION AND OBJECTIVES Motivation and objectives 26 Chapter 2 Motivation The wide range of applications that Polystyrene (PS) plastics varieties find in the daily life, makes that its consumption grows continuously. Nevertheless, due to its resistance to biodegradation, PS wastes have become one of the most important environmental issues. By this reason, polymer wastes needs to be dealt with to preserve the natural resources, decrease the hazardous impacts, reduce the amount of solid waste ending up in landfills, and enhance the sustainability for future generations. However, hitherto recycling of the mentioned polymer wastes is not economically feasible since high costs are implied. The oversupply of virgin PS is the cause of its low price, moreover, the low density of PS means high volume of wastes and consequently, major efforts are needed to offset the costs of transport. Also, the low quality of the recycled product obtained from the traditional methodologies promotes the lack of a market for the recovered PS. In view of the aforementioned environmental and economical issues related to the production of Polystyrene wastes, new processes should be designed with the aim of reducing the amount of PS wastes, achieving a high-quality recovered product and succeeding an economically feasible process. Figure 2.1 shows the scheme of the suggested process. Terpene Oil/CO 2 PS/Terpene Oil/CO2 Polystyrene Terpene Solvents CO2 PS/Terpene Oil Dissolution Figure 2.1. Schematic diagram for the recycling of Polystyrene wastes to produce high-added value final product by an economically feasible process. With regard to the Figure 2.1, the general process for the recycling of Polystyrene wastes can be divided in two main steps: the dissolution of Polystyrene wastes in 27 Motivation and objectives terpene oils and the separation of the polymer from the solution by CO2 at high pressure. The dissolution of wastes in terpenes allows the effectively decrease of the PS volume, since the density of the oils is higher than those of PS wastes and volume could be reduced between 5 and 15 times depending on the concentration. Next, the CO2 solubilises the terpene while the polymer remains changeless except that CO2 is absorbed among the polymeric chains and is able to tune the polymer structure. Finally, Polystyrene is recovered as microcellular foam, a very high-added value product. Objectives According to the environmental issues concerning the Polystyrene wastes, the aim of this Thesis is their recycling by means of an environmentally friendly technology based in terpene oils and supercritical carbon dioxide. The traditional recycling methods entail ageing and degradation of the polymer (mechanical recycling), environmental contaminations (pyrolysis) or even polluting praxis. In this work, a new technology has been proposed for the recycling process of Polystyrene wastes with the aim of achieving microcellular foams according to the Green Chemistry principles. In order to achieve this general aim, the research work has been structured in the following particular objectives: 28 Selection of the most suitable terpene solvents to perform the dissolution of Polystyrene wastes. Study of the influence of temperature, molecular weight and source on the solubility results. Determination of the Vapour-Liquid Equilibrium data of the systems terpene oils/CO2. Study of the effect of pressure and temperature on the sorption of CO 2 in PS. Knowledge of the phases equilibrium of the ternary mixtures CO2/Limonene/PS and CO2/p-Cymene/PS. Limitation of the regions where the separation could be achieved. Analysis of the modification of properties under high pressure of the mentioned ternary systems. Study of viscosity, glass transition temperature and interfacial tension. Establishment of the most suitable conditions to perform the precipitation and foaming processes for the recycling of Polystyrene wastes. Preparation, characterization and optimization of the foaming process. Life Cycle and economic assessment of the global process for the recycling of PS wastes by dissolution in terpene oils and CO2. Chapter 2 Work plan According to the objectives and the main goal of this research vocation, the next work plan was developed. Recycling of Polystyrene wastes by dissolution in terpene oils and scCO2 Binary Systems Polystyrene/ Terpene oils Terpene oils/ scCO2 Ternary System Influence of Pressure Influence of Temperature Polystyrene/ scCO2 Equilibrium of phases Glass transition temperature Properties Influence of Concentration Foaming process Study of Pressure, temperature, concentration, contact time and depressurization time Interfacial Tension Viscosity Characterization: pore size distribution, Tdeg, Tg, molecular weight distribution, amount of terpene Optimization Figure 2.2. Work plan developed along the thesis. 29 Motivation and objectives Thesis Overview Chapter 1, shows a general review about the plastic wastes management and recycling options. Nowadays, many different techniques are being studied since the optimum process for the recovery of high-added value products from plastic waste streams is still a challenge. In Chapter 2, the motivation and objectives of the research study are shown in order to provide the backbone of the thesis. Next, Chapter 3 details the most relevant experimental setups and procedures to perform the experiments whose results are shown in the subsequent sections. The results have been divided in five main chapters: the binary and the ternary systems, the study of their properties, the foaming process for the recycling of PS wastes and the Life Cycle and economic Assessment. Chapter 4 is focused on the study of the binary systems: Polystyrene/Terpene oils, Terpene oils/CO 2 and Polystyrene/CO2. Also, several properties of the mentioned systems were studied; particularly, glass transition temperature (Tg) and viscosity (). This issue has been part of the next papers: A practical approximation to design a process for polymers recycling by dissolution, in the journal Afinidad; The selective dissolution technique as initial step for polystyrene recycling, in the journal Waste and Biomass Valorization; Modeling the phase behaviour of essential oils in supercritical CO 2, in the journal Industrial & Engineering Chemistry Research and Modification of Polystyrene properties by CO2: experimental study and correlation in the journal Chemical Engineering Science. Although it could be thought that once binary systems are defined, the design of the process would be carried out, the determination of the mixtures and the evaluation of their properties is crucial to succeed the global process. In Chapter 5, two ternary systems (CO2/Limonene/PS and CO2/p-Cymene/PS) were selected to study their phases equilibrium. Part of the research concerning the ternary systems and their properties has published in the following papers: Highpressure phase equilibria of Polystyrene dissolutions in Limonene in presence of CO 2 and Determination of the high-pressure phase equilibria of Polystyrene/p-Cymene in presence of CO2 both in the Journal of Supercritical Fluids. Chapter 6: and their properties under high pressure. In this case, glass transition temperature (Tg), viscosity () and interfacial tension (IFT) at high pressure were studied. The most relevant results have been published in the papers: The effect of CO2 on the viscosity of polystyrene/limonene solutions and Development of a strategy for the foaming of polystyrene dissolutions in scCO2, all of them in the Journal of Supercritical Fluids. Chapter 7 addresses the precipitation and foaming process of the ternary systems previously studied. Initially, the feasibility of the separation process was achieved and next, the study of the variables that influenced on the process was carried out. The optimization of the foaming process was performed in order to achieve microcellular foams, it means, minimize the average cells size and its standard 30 Chapter 2 deviation, while the cells density is maximized. The first results were shown in the paper entitled Development of a strategy for the foaming of polystyrene dissolutions in scCO2, in the Journal of Supercritical Fluids and Foaming process from Polystyrene/p-Cymene solutions using CO2 in Chemical & Engineering Technology. The further study of the foaming process was presented under the title Preparation and characterization of polystyrene foams from limonene solutions, in the Journal of Supercritical Fluids. Finally, in Chapter 8, the Life Cycle and Economic Assessment of the recycling of PS wastes using terpene oils and CO2 as antisolvent was performed. 31 Motivation and objectives 32 Chapter 3 MATERIALS AND EXPERIMENTAL METHODS Experimental Chapter has been divided in four different sections: materials, experimental setups, procedures and characterisation. A brief summary of the most relevant equipments used in this research work are shown below. Materials and Experimental Methods 34 Chapter 3 3.1. Materials Polymers: Polystyrene, Molecular weight 280000 g/mol. CAS # 9003-53-6. Supplied by Sigma-Aldrich. Polystyrene, Molecular weight 192000 g/mol. CAS # 9003-53-6. Supplied by Sigma-Aldrich. Polystyrene, Molecular weight 350000 g/mol. CAS # 9003-53-6. Supplied by Sigma-Aldrich. Expanded Polystyrene (EPS), Molecular weight: 79309, Polydispersity inde: 1.791. Waste from packaging. Extruded Polystyrene (XPS), Molecular weight: 89068, IP: 1.845. Supplied by Tecnove Fiberglass. The composition of the XPS wastes is detailed in Table 3.1. Table 3.1. Extruded polystyrene (XPS) waste composition. Composition % w/w XPS Flame retardant (HBCD) Nucleating agent (Talc) Ethyl chloride Pigments 90-93 2-3 0-1 0-2 0.2-0.3 Solvents: Carbon Dioxide, purity 99.8 %. CAS # 124-38-9. Supplied by Praxair. Anisole, purity 99%. CAS # 100-66-3. Supplied by Sigma-Aldrich. Cinnamaldehyde, purity 98%. CAS # 104-55-2. Supplied by Sigma-Aldrich. p-Cymene, purity 99%. CAS # 99-87-6. Supplied by Sigma-Aldrich. Eucalyptol, purity 98%. CAS # 470-82-6. Supplied by Sigma-Aldrich. Geraniol, purity 97%. CAS # 106-24-1. Supplied by Sigma-Aldrich. Limonene, purity 97%. CAS # 138-86-3. Supplied by Sigma-Aldrich. Linalool, purity 95%. CAS # 78-70-6. Supplied by Sigma-Aldrich. -Phellandrene, purity 95%. CAS # 99-83-2. Supplied by Sigma-Aldrich. -Pinene, purity ≥ 95%. CAS # 80-56-8. Supplied by MERCK. -Terpinene, purity 95%. CAS # 99-85-4. Supplied by Panreac. -Terpineol, purity 98%. CAS # 98-55-5. Supplied by Sigma-Aldrich. 35 Materials and Experimental Methods Analytical reagants Gel Permeation Chromatography: Tetrahydrofuran, (THF), CHROMASOLV Plus, purity ≥ 99.9%. CAS # 10999-9. Supplied by Sigma-Aldrich. Polystyrene Standard for calibration with molecular weight ranged from 370 and 177000, supplied by Waters. 3.2. Experimental setups 3.2.1. Solubility determination at atmospheric pressure The solubility of Polystyrene in terpene oils was determined by saturation of the solvents at atmospheric pressure, room temperature (25 ºC) and stirring at 1000 rpm to promote dissolution and homogenization process. The experimental setup is shown in Figure 3.1. Figure 3.1. Experimental setup used for the determination of the solubility of Polystyrene in terpene oils. 5 g of PS pellets or wastes were placed in glass test tubes with 5 ml of terpene solvents. The tubes were sealed with Parafilm ® to prevent evaporation of solvents. 3.2.2. Variable-Volume High Pressure View Cell Experimental measurements of the high-pressure phase equilibria were carried out using a variable-volume cell model ProVis 500 (from Eurotechnica, Hamburg, Germany), as it is depicted in Figure 3.2 and Figure 3.3. Figure 3.2 shows the experimental high-pressure view cell used for the determination of vapour-liquid equilibrium behaviour of binary mixtures CO2/terpene oils. 36 Chapter 3 T A Pressure Generator PI C B G-1 TIC G-2 F D F-1 Gas sample outlet Liquid sample outlet F-2 CO2 inlet Figure 3.2. Experimental setup employed for high-pressure phase equilibrium in the binary systems. PG: pressure generator; PI-1: manometer; PI-2: pressure digital indicator; T: liquid supply tank; TI: temperature digital controller. The equipment shown in Figure 3.2. consists of a variable-volume cell, supplied with a front and upper sapphire window and light for visual observation of phase separation. The cell has a maximum capacity of 50 cm3 and contains a piston system, consisting in a manual pressure generator, a cylinder and a movable piston made of Teflon to avoid pressure drops (Valves A and C) only when samples are taken. The movable piston separates the equilibrium chamber from the pressurizing circuit. To allow a smooth displacement of the piston inside the cylinder, the piston was driven by a manual pressure generator and water was used as the pressurizing fluid. All the system is heated externally by an air bath made of Poly(methyl methacrylate) capable to resist temperatures up to 80ºC being the temperature inside the equilibrium chamber measured by a thermocouple (PT100) coupled to a digital led model testo 925 (Lenzkirch, Germany). Figure 3.3 shows a schematic diagram of the high-pressure view cell modified for the measurement of phase behaviour in ternary mixtures containing CO2/terpene oils/ Polystyrene. 37 Materials and Experimental Methods T A Pressure Generator PI C B TIC D G-1 G-2 Gas sample outlet CO2 inlet Figure 3.3. Experimental setup employed for high-pressure phase equilibrium in the ternary systems. PG: pressure generator; PI-1: manometer; PI-2: pressure digital indicator; T: liquid supply tank; TI: temperature digital controller. According to Figure 3.3, a loop of 20 cm3 was introduced into the high-pressure view cell to promote the homogenisation of the ternary mixtures. The gear pump (Micropump® SerieGAH) provides appropriate stirring inside the cell by the recirculation of the gas. The rest of the setup was not modified, although in the determination of the equilibrium of ternary mixtures only gas samples were withdrawn. 3.2.4. High-Pressure Differential Scanning Calorimetry Differential Scanning Calorimetry (DSC) was used to determine the influence of CO2 on the glass transition temperature (Tg) of Polystyrene and its solution in terpenes at high pressure. A schematic diagram of the experimental setup is shown in Figure 3.4. 38 Chapter 3 CO2 inlet V-3 PI C-1 V-2 V-1 E-1 V-4 V-5 F-1 C-2 P-1 Water P-2 CO2 out V-6 I-1 Figure 3.4. Schematic diagram of the experimental setup for the measurement of the glass transition temperature. V-1, V-2: check valves; E-1: cooler; V-3: purge valve; F-1: filter; P-1: membrane pump; P-2: syringe pump; V-4, V-6: regulator valves; V-5: micrometric valve; PI: digital pressure gauge; C-1, C-2: crucibles for sample and reference, respectively; I-1: turbine flow meter. The experiments were performed in a SENSYS evo DSC (Setaram, Madrid, Spain) equipped with two high pressure Inconel crucibles (C-1 and C-2) enabled measurements up to 400 bar. 3.2.5. Rheometer A Physica MCR 301 rheometer (Anton Paar, Hertfor, UK) was used to determine the effect of pressure and temperature on the viscosity of Polystyrene. 7 8 9 . 4 5 6 0 1 2 3 Figure 3.5. Schematic diagram of the experimental setup for the measurement of the viscosity of Polystyrene at high pressure. 39 Materials and Experimental Methods The rheometer is equipped with a pressure cell which allows to work at high pressure, a synchronous motor with air bearing system to stabilize the measurement and in our case parallel-plate geometry was used to decrease the error in gap size. A peltier temperature control allows to get an uniform temperature in the sample up to 200°C. Data were analyzed by RheoPlus software. 3.2.6. Quartz Viscometer The effect of pressure, temperature and PS concentration on the solution viscosity was determined using a Quartz Viscometer (Flucon Fluid Control GmbH, Germany). A schematic diagram is shown in Figure 3.6. The most important element of the quartz viscometer is a piezoelectric torsion oscillator of SiO 2, which is fed with alternating voltage of high-frequency and is placed inside the high pressure vessel. The quartz sensor is located inside a thermo-stated pressure cell and includes an electric thermostat using aluminium alloy blocks as the medium of heat transfer. The temperature is controlled by means of an electrically heated thermostat jacket with a temperature range between 20 and 200 ºC. A thermocouple inside the high pressure cell allows monitoring and recording of the temperature. The viscometer is connected to a controlling computer in which the software (QVis) is installed to allow automation of measurements and the subsequent evaluation of experimental data. TI PI (a) CO2 inlet CO2 outlet (b) VI TIC Figure 3.6. Schematic diagram of the high-pressure quartz viscometer experimental setup (a) high pressure stainless steel vessel; (b) quartz sensor. The implementation connected to the viscometer is PI: pressure transducer; TI: temperature digital sensor; TIC: temperature digital controller; VI: viscosity indicator. 40 Chapter 3 3.2.3. Pendant droplet High Pressure View Cell The pendant droplet method, introduced by Andreas and Hauser [1], was the selected to determine interfacial tension between PS dissolution and CO 2, because it has proved to be one of the most practicable methods for systems under high pressure. The experiments were also carried out in a high-pressure view-cell model PD-E1700 (Eurotechnica GmbH, Hamburg, Germany) containing sapphire windows to observe the pendant drop, schematic diagram is shown in Figure 3.7. and it has been previously referenced [2-4]. The equipment can be applied at pressures up to 500 bar and up to 200 ºC. Solution Tank Pressure Generator PI PI CO2 inlet Figure 3.7. Experimental setup with a view chamber where is placed the pendant droplet. The shape of the drop depends on the interfacial tension and on the density difference of the surrounding fluid phases. Drop shapes were recorded by a commercial CCD video technique, and later, the images were analyzed by image processing software called drop shape analysis (DSA) provided by Krüss GmbH. Density difference between the coexisting fluids (Polystyrene solution and the surrounding CO2) were calculated to determine interfacial tension according to Laplace equation of curved fluid surfaces [3]. 3.2.7. Foaming setup Foams were prepared in a home-made batch-type device consisted on a 316 stainless steel high-pressure vessel (350 ml) as it is depicted in Figure 3.8. 41 Materials and Experimental Methods BPR CO2 inlet V-2 PI V-3 V-4 V-5 V-6 I-1 V-1 F-1 TIC Figure 3.8. Schematic diagram of the experimental foaming rig. V-1, V-3: check valves; V-2: purge valve; E-1: cooling system; F-1: filter; BPR: back-pressure regulator; P-1: pump; V-4, V5, V-6: regulator valves; PI: manometer; C-1: foaming vessel; I-1: turbine flow meter; TIC: temperature digital controller. Green lines represent the refrigerated pipes. As Figure 3.8 shows, the foaming rig is divided in three modules: Pressurization module: where CO2 from the tank is cooled and pumped up to the operating pressure by a dosing pump (Milton Roy-Milroyal D, France). A purge valve allows the checking of the CO2 phase before the pumping. Main module: the high-pressure stainless steel vessel was heated by an electric coat digitally controlled. Venting module: consists of several regulator valves which allows the control of the depressurization flow. 3.2.8. Supercritical Antisolvent precipitation setup The antisolvent precipitation of Polystyrene solutions by CO 2 was performed in a setup similar to the described in the previous section. Figure 3.9 shows the experimental setup. 42 Chapter 3 V-2 PI BPR V-3 V-4 PI V-6 V-5 V-1 TIC V-7 V-8 Figure 3.9. Schematic diagram of the experimental setup for the precipitation of Polystyrene from its solution. V-1, V-3: check valves; V-2: purge valve; V-4, V-5, V-6, V-7, V-8: regulator valves; PI: manometer; TIC: temperature digital controller. Green lines represent the refrigerated pipes and red lines show the heated pipes to balance Joule-Thomson effect. The general scheme is similar to the described in the section 3.2.7, where three modules are distinguished. But in this case, the pressurization module is divided in two parts: CO2 and solution fed. Polystyrene solution is compressed and fed from a burette by mean of a dosing pump (Milton Roy-Milroyal D, France) through a stainless steel nozzle of variable diameter (1/16 inch or 500 nm) located at the top of the high-pressure vessel. Also, the depressurization module is modified, the CO 2 and the soluble terpene are removed from the vessel through and outlet point at the bottom and expanded into the liquid solvent recovery vessel. This expansion and the flow were regulated by a micrometering valve and a bubble flowmeter. 3.3. Experimental Procedures 3.3.1. Determination of the solubility of Polystyrene in Terpene oils Solubility determination was carried out by thermogravimetric analysis (TA Instruments-SDT 2960) at a heating rate of 10ºC/min from room temperature to 600ºC under a nitrogen atmosphere. Once the samples were saturated (around 48 43 Materials and Experimental Methods hours) two samples were withdrawn from the clear saturated solution to minimize experimental errors. Thermogram analysis obtained, showed two weight losses, the first one belongs to the solvent boiling (175ºC) and the second one to Polystyrene decomposition (421ºC). 3.3.2. High Pressure Vapour-Liquid Equilibria of Binary mixtures Terpenes/CO2 The experimental setup used to determine the Vapour-Liquid Equilibria of binary mixtures consisting of terpene and CO 2 was shown in Figure 3.2. Terpenes (pCymene, Limonene and -Terpinene) were introduced into the cell that was purged with CO2 at low pressure to remove the residual air. Then CO 2 was allowed to flow into the cell (Valve D) and operating pressure and temperature were reached. Once the temperature was stabilized and the desired pressure was achieved, the magnetic stirrer was turned on. To confirm that equilibrium conditions were reached, previous experiments were carried out taking samples small enough not to affect the global composition of system at different times. When the solvent concentration in CO 2 remained constant between two samples in the less favourable conditions (low temperature and pressure), the time was selected to be sure that the equilibrium was reached and withdraws the following samples. Once the equilibrium conditions were reached, stirring was stopped, and the mixture was allowed to repose in order to favour the separation of the phases. Samples from the top (Valves F and F1) and the bottom (Valves G1 and G2) of the equilibrium cell were withdrawn using capillary lines and needle valves and decompressed to atmospheric pressure. Valves were thermostatized at the same temperature of the equilibrium cell to avoid condensation. The manual pressure generator was employed to keep pressure constant during sampling (± 0.01 MPa). For the determination of the amount of CO2 and terpene in each phase, two samples were withdrawn from the bottom and the top of the equilibrium cell and expanded into a glass vial to check the reproducibility. The vials were weighted in a precision analytical balance with 0.0001 g accuracy to determine the amount of terpene oil collected. The amount of CO2 was measured during the sampling through a gas meter. 3.3.3. High Pressure Phase CO2/Terpenes/Polystyrene Behaviour of Ternary mixtures The experimental equipment used to determine the high pressure phase equilibrium of ternary mixtures was shown in Figure 3.3. Polystyrene and Terpene oil solutions 44 Chapter 3 were prepared in glass vials and introduced into the cell which was purged with CO2 at low pressure to remove the residual air. Then CO 2 was allowed to flow into the cell (by valve D). Once the temperature was stabilized, the desired pressure was achieved. After that, the gear pump was stopped, and the mixture was allowed to repose in order that phase segregation occurred. As it was explained in the preceding experimental procedure, the time to reach equilibrium conditions was previously determined. Samples from the top of the equilibrium cell were withdrawn isobarically through a six port valve connected to a 20 cm3 loop using capillary lines and needle valves and decompressed to atmospheric pressure (by Valves F and F1). Valves were thermostatized at the same temperature of the equilibrium cell. The manual pressure generator was employed to keep pressure constant during sampling (± 0.1 MPa). For the determination of the amount of solvent solubilised in CO 2, two samples were withdrawn from the equilibrium cell and expanded into a glass vial, which was weighted before and after sampling in a precision analytical balance with 0.0001 g accuracy. Limonene was collected in glass vials and separated from the CO2 by using a trap. The amount of CO2 was measured through a gas meter. The CO2 density was calculated as a function of pressure and temperature with the equation of Bender [5]. 3.3.4. Glass transition temperature at high pressure The influence of CO2 on the glass transition (Tg) of Polystyrene and its solution in terpenes was determined using the experimental setup shown in Figure 3.4. Initially, the samples (the polymer or its solutions) were placed in the crucibles, weighted and filled with CO2. Next, CO2 was cooled (in E-1) and pressurized to the desired level by means of a positive-displacement pump (P-1) and a syringe pump (P-2) which controlled the feed amount of gas. After that, samples were annealed at the desired pressure for 24 hours to ensure the total sorption of CO2. Differential Scanning Calorimetry scans were made using an initial heating at 25 ºC/min up to 120 ºC to release thermal and absorption history of Polystyrene and to provide a better fit of the polymer in the crucible. The samples were then cooled at the same rate and annealing for 30 minutes. Tg measurements were carried out during the second heating, and it was identified from the change in heat flow resulting from a change in heat capacity at the transition temperature during each scan. Once the measurements were performed, the vessel was vented by opening a discharge valve (V-6) and the amount of CO2 was measured by a turbine flow meter (I-1). 45 Materials and Experimental Methods 3.3.5. Determination of the viscosity of Polystyrene The experimental measurement of the viscosity of Polystyrene at high pressure was carried out using the rheometer described in Figure 3.5. The Polystyrene was milled to fill the surface of the cell that was sealed and charged with CO2 using a syringe pump. The samples where left for 15 min to allow the diffusion of CO2 into the samples. The shear rate was ranged from 0.001 to 1000 min-1 over 5 min and data points were recorded every 3 seconds. At least three analyses were run for the different conditions. 3.3.6. Determination of the viscosity of Polystyrene solutions The study of the influence of pressure, temperature and concentration on the viscosity of Polystyrene/Terpene oil solutions in the presence of CO 2 was performed using the quartz viscometer depicted in Figure 3.6. The experimental procedure requires the Polystyrene/terpene solution to be placed inside the autoclave which was then filled with CO2. The pressure was measured by a pressure sensor and was kept constant during each measurement. The major requirement for obtaining correct results when measuring was to ensure the complete immersion of the quartz into the solution and to ensure the absence of bubbles in the fluid mixture. This was achieved by observing the solutions under the same pressures in a view cell. In all cases, the samples were left for 5 minutes to allow the CO 2 to diffuse and homogenise into the polymer solution. A series of standard oils covering a wide range of viscosities from 0.11 to 10000 cP were used as comparison fluids with known Newtonian behaviour. The calibration of the quartz viscometer was carried out in a range of temperature from 20 to 100 ºC, using a mineral oil with low viscosity (D90) and with high viscosity (D7500); the average standard deviation obtained among the measurements was ± 1.36 mPa·s and 174.56 mPa·s, respectively. 3.3.7. Determination of high-pressure Interfacial Tension The measurement of the Interfacial Tension (IFT) of Polystyrene/terpene oil solutions at high pressure were carried out using the experimental setup shown in Figure 3.7. The experimental procedure consisted of bringing to the working temperature the viewing chamber, and then introducing CO2 to the desired pressure. Next, the drop liquid is generated at the capillary tip using a hydraulic piston system. When a stable droplet is obtained, it is recorded by the CCD-unit. Once the droplet shape is digitalized, the software finds the solution that best fits the experimental drop profile by minimizing the deviation between experimental 46 Chapter 3 and calculated. The result is the value of the IFT at the desired temperature and pressure. Drop age was considered to assure that the equilibrium value of the interfacial tension was kept constant along time (Figure 3.10). 40 35 30 (mN/m) 25 20 15 10 PS/Cymene C0=0.05 g/ml 5 0 bar 50 bar 0 0 100 200 300 400 500 600 Time (s) Figure 3.10. Influence of time on the Interfacial tension of Polystyrene/p-Cymene solution at 30 ºC. As it is observed in Figure 3.10, interfacial tension is reached relatively fast as a result of a rapid achievement of the equilibrium composition between the drop and the inter-phase. Since the real mixture densities that resulted from mutual solubility are not known, the experimental density values for the pure components as a function of temperature and pressure were used. In case the density must be corrected after the experiments, the old IFT directly obtained must be divided by the old density difference. The resulting density-related value is then multiplied by the new density difference. 3.3.8. Foaming of Polystyrene at high pressure The foaming of Polystyrene from its solution at high pressure was performed using the experimental setup described in Figure 3.8. Initially, the polymer solution was placed in the vessel and it was slightly heated. Next, liquid CO 2 from a stainlesssteel cylinder was cooled (E-1), filtered (F-1), and compressed by a positivedisplacement pump (P-1). The lines were cooled to avoid cavitation of the pump. The pressure was regulated by a back-pressure regulator (BPR) and checked by a manometer (PI) before passed to a 350 mL stainless-steel cylinder (C-1) through a regulation valve (V-4). Maximum operating conditions of the equipment were 120 ºC and 330 bar. The CO2 stayed inside the vessel during the selected contact time in 47 Materials and Experimental Methods order to assure that the saturation conditions were reached. The temperature was kept constant through a digital controller (TIC) which regulated the electric current by means of a resistance placed around the vessel (± 0.1 ºC). Temperature and pressure were kept constant during the time of the experiment. At the end of this period, the vessel was depressurized opening a discharge and a regulator valve (V-5 and V-6, respectively) that was controlled manually by the measurement of the flow in a turbine flow meter (I-1) in order to induce and control the cell nucleation. Once the pressure was completely released, the resistance was got away and the highpressure vessel was allowed to cool, and quickly opened up to take out the foamed samples for subsequent analyses. 3.3.9. Precipitation of Polystyrene by Supercritical Antisolvent process The precipitation of Polystyrene from its solution using CO 2 as antisolvent was performed using the experimental setup described in Figure 3.9. Initially, the highpressure vessel was heated or cooled to the operating temperature. The temperature was kept constant through a digital controller (TIC) which regulated the electric current by means of a resistance placed around the vessel (± 0.1 ºC). Next, the CO2 was cooled, pressurized and fed to the vessel at the working pressure. The pressure was regulated by a back-pressure regulator (BPR) and checked by a manometer (PI). On the other hand, polymer solution was fed to the pump to be pressurized up to the operating pressure controlling the flow. Once the conditions (temperature, pressure and flow) were reached, the CO2 together with the soluble terpene were removed from the vessel by opening the heated valves V-7 and V-8. The outflow was controlled manually by mean of a bubble flowmeter and terpenes were condensed in the liquid solvent recovery glass vessel. Temperature, pressure and flows were kept constant during the time of the experiment. When the solution was completely fed to the reactor, valves V-5 and V-6 were closed and CO2 was flushed to the reactor through the bottom pipe. 3.4. Characterisation 3.4.1. Thermogravimetric Analysis (TGA) Thermogravimetric analysis (TA Instruments-SDT 2960) was used to: 48 determine the composition of solution (Polystyrene and terpene) after equilibrium conditions at high pressure were reached. quantify the amount of residual solvent absorbed on the particles, fibres or foams. Chapter 3 calculate the degradation temperature of the Polystyrene. Figure 3.11. Thermogravimetric balance. The standard procedure consisted of heat the samples (3-10 mg) at aheating rate of 10ºC/min from room temperature to 600ºC under a nitrogen atmosphere. 3.4.2. Thermogravimetric and Mass Spectrometer Analysis (TGMSA) The analysis of the gas products from the pyrolysis of Polystyrene was carried out in a thermogravimetric analyser (TGA-DSC 1; METTLER TOLEDO) coupled to a mass spectrometer (Thermostar-GSD 320/quadrupole mass analyser; PFEIFFER VACUUM) with an electron ionization voltage at 70 eV (Figure 3.12). Figure 3.12. Thermogravimetric and Mass Spectrometer Analysis The samples were heated from room temperature up to 600 ºC at a heating rate of 10 °C/min. The experiment was performed under argon atmosphere. In order to identify ions with m/z in the range 0–175, a preliminary broad scan was performed at a heating rate of 10 ºC/min. Each sample was analyzed at least three times, and the average value was recorded. The experimental error of these measurements was calculated, obtaining an error for all studied samples of ±0.5% in weight loss measurement and ±2 ºC in temperature measurement. 49 Materials and Experimental Methods 3.4.2. Differential Scanning Calorimetry Analysis (DSC) The differential scanning calorimetry (DSC) was used to determine the glass transition temperature of the recovered Polystyrene. The apparatus used in the analysis is presented in Figure 3.12. Figure 3.12. Differential scanning calorimeter. It consisted of a calorimeter DSC Q1000 (TA Instruments) equipped with refrigerated cooling system (RCS) and autosampler. Typical sample weights were 410 mg. All the measurements were run in aluminium hermetic pans under a nitrogen flow of 50 mL/min. The software used in the analysis was TA Universal Analysis 2000. The experimental run consisted of a first heating up to 120 ºC, a cooling up to 0 ºC and a second heating where the glass transition temperature was determined. During the three stages, the heating or cooling rate was 10 ºC/min. 3.4.3. Optical Microscopy (OM) Polymer foams, microparticles and fibres were analyzed and imaged using optical microscopy in order to determine their pore size and morphology. It is a Carl Zeiss, model Axio Imager A1 10x and 40x (Figure 3.13), equipped with a camera Nikon Digital Sight model DS-2Mv which did the microphotographs. The software used to process the data was Nis Elements BR. 50 Chapter 3 Figure 3.13. Optical microscopy. A Nikon Eclipse E 200 microscope (Kingston, England) equipped with Nikon Application Suite Interactive Measurement software was used to analyze the polymer foams. The optical microscopy technique provides images of the foams with good resolution and this allows study of the morphology of the recovered polymer. 3.4.4. Scanning Electron Microscopy (SEM) Scanning electron microscope is an extremely useful tool for the study of the surface of the samples because it offers a better resolution than the optical microscope. The surface features of the Polystyrene foams were observed by using a Quanta 250 (FEI Company) for scan electron microscopy (SEM) (Figure 3.14). Figure 3.14. Scanning Electron Microscopy. 51 Materials and Experimental Methods 3.4.5. Determination of Molecular Weight Polymer average molecular weight and its distribution were determined using gel permeation chromatography (GPC). This technique is included in the size exclusion chromatography methods and it is the most powerful and widely used in this field. The average molecular weights resulted from this analysis are in weight (M w), in number (Mn) and the polydispersity index (PDI). The apparatus used for this characterization, shown in Figure 3.15, was a chromatograph (Waters) equipped with two different columns, Styragel HR1 and Styragel HR4 whose molecular weight interval covered from 100 to 500 000 g/mol. It is controlled by a computer system supplied with the Breeze software which controls the entire process but also stores and processes the results. Figure 3.15. Gel permeation chromatograph. The analysis conditions are shown in Table 3.15. The polymer solutions were also filtered before the measurements using a syringe with an attached filter of a pore size of 0.45 µm. The chromatograph was calibrated using the kit of PS standards described in Table 3.2. Table 3.2. Analysis conditions in molecular weight measurements. Parameter Solvent Temperature Injection volume Sample concentration Flow rate 52 Values THF 35 ºC 100 µL 1.5 mg/mL 1 mL/min Chapter 3 References [1] J.M. Andreas, E.A. Hauser, W.B. Tucker, Boundary tension by pendant drops, Journal of Physical Chemistry, 42 (1938) 1001-1019. [2] P.T. Jaeger, M.B. Alotaibi, H.A. Nasr-El-Din, Influence of compressed carbon dioxide on the capillarity of the gas-crude oil-reservoir water system, Journal of Chemical and Engineering Data, 55 (2010) 5246-5251. [3] P.T. Jaeger, R. Eggers, H. Baumgartl, Interfacial properties of high viscous liquids in a supercritical carbon dioxide atmosphere, Journal of Supercritical Fluids, 24 (2002) 203-217. [4] P.T. Jaeger, J.V. Schnitzler, R. Eggers, Interfacial tension of fluid systems considering the nonstationary case with respect to mass transfer, Chemical Engineering and Technology, 19 (1996) 197-202. [5] E. Bender, Equations of state for ethylene and propylene, Cryogenics, 15 (1975) 667-673. 53 Materials and Experimental Methods 54 Chapter 4 THE BINARY SYSTEMS In this Chapter the binary systems involved in the recycling of Polystyrene were studied individually. Since the recycling process is planned, the main idea is to maximize the amount of Polystyrene in the solutions. By this reason, a first screening among the most common terpene oils was studied. Next, the selected solvents must be soluble in CO2 to recover the terpene and precipitate the polymer. The Vapour-Liquid Equilibrium (VLE) of the binary mixture CO 2/terpene oils was determined at mild temperatures. Finally, the influence of CO2 on the properties of Polystyrene was investigated as initial approach of the global process. Binary Systems 56 Chapter 4 Based on the papers: C. Gutiérrez, M.T. García, I. Gracia, A. De Lucas, J.F. Rodríguez, A practical approximation to design a process for polymers recycling by dissolution, Una aproximación práctica para el diseño de un proceso de reciclado de polímeros mediante disolución, Afinidad, 68 (2011) 181-188. C. Gutiérrez, M.T. García, I. Gracia, A. De Lucas, J.F. Rodríguez, The selective dissolution technique as initial step for polystyrene recycling, Waste and Biomass Valorization, 4 (2013) 29-36. C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. de Lucas, M.T. García, Modeling the phase behavior of essential oils in supercritical CO2, sent to Industrial & Engineering Chemistry Research. C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. de Lucas, M.T. García, Modification of Polystyrene properties by CO2: experimental study and correlation, sent to Chemical Engineering Science. Graphical abstract 140 T: 25ºC T: 30ºC T: 40ºC Experimental data at 30ºC Experimental data at 40ºC 120 Pressure (bar) 100 80 60 40 20 0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.98 0.99 1.00 wCO Liq/Vap 2 Terpene oils/CO2 Polystyrene/ Terpene oils CO2/Polystyrene 0.20 0.15 0.10 This work at T: 30 ºC Density of CO2 at T: 30 ºC 1200 1000 0.15 800 0.10 600 400 0.05 3 0.05 200 Anisole Linalool g-Terpinene Geraniol Limonene Eucalyptol a-Pinene a-Phellandrene Cynammaldehyde p-Cymene 0.00 a-Terpineol Solubility (g/ml) 0.25 0.20 Sorption (gCO 2/gPS) 0.30 Density of CO 2 (kg/m ) Recycling of Polystyrene wastes 0.35 0.00 0 57 Binary Systems RESUMEN El Poliestireno (PS) es uno de los polímeros más comúnmente utilizados debido a su gran versatilidad, lo que genera una gran cantidad de residuos que es necesario gestionar de manera adecuada. En este trabajo, se ha propuesto un nuevo proceso para su reciclado que consiste en la disolución de los residuos en aceites terpénicos y su posterior separación mediante CO2, ya que éste es capaz de solubilizar a los componentes terpénicos, pero no al polímero. En este Capítulo, se muestra el estudio pormenorizado de cada uno de los sistemas binarios implicados en el proceso. En primer lugar, se estudió la solubilidad del PS en los aceites terpénicos seleccionados. En este primer apartado, se determinó tanto teórica, como experimentalmente, la influencia del peso molecular del polímero, la temperatura a la que se lleva a cabo la disolución y el origen de los residuos en la solubilidad del polímero. En segundo lugar, se estudió el equilibrio líquido-vapor entre los aceites terpénicos y el CO2. Para ello, se llevó a cabo una recopilación bibliográfica que permitiera generalizar el comportamiento de éstos, para a continuación aplicar modelos semiempíricos (Chrastil) o teóricos (Ecuación de estado de Peng-Robinson). Por último, se analizó el efecto del CO2 sobre el Poliestireno, ya que su adsorción entre las cadenas poliméricas, es la responsable del hinchamiento y la plastificación del polímero. Debido a este fenómeno, se producen cambios en las propiedades térmicas y físicas del Poliestireno que repercuten en una disminución de su temperatura de transición vítrea, su tensión superficial o su viscosidad. De acuerdo con esto, en este primer Capítulo de resultados, se pretenden sentar las bases que permitirán conocer mejor el proceso global de reciclado de residuos de Poliestireno. 58 Chapter 4 ABSTRACT Polystyrene (PS) is widely used in different applications such as packaging or insulation materials and consequently enormous volume of wastes are generated. In this work, a new recycling process has been proposed, consisting of the dissolution of Polystyrene wastes in terpene oils and its separation using CO2 as solvent of the terpenes and antisolvent of the polymer. In this Chapter, a detailed study of each binary system is shown. Firstly, the solubility of PS in some selected terpenes was studied. In this section, the influence of the molecular weight of the polymer, the temperature of the solutions and the waste sources was analyzed theoretically and experimentally over the solubility of the polymer. Secondly, the vapour-liquid equilibrium (VLE) of terpene oils in CO2 was determined. A wide literature review was accomplished with the aim to generalize a procedure which estimates general trends of terpenes in CO 2. Semi-empirical (Chrastil’s equation) and theoretical models (Peng-Robinson Equation of State) were applied to correlate and predict VLE data. Finally, the interactions between the CO 2 and Polystyrene were analyzed. The sorption of CO2 into the polymer causes its swelling and plasticisation which modifies the physical and thermal properties. The decrease of glass transition temperature, interfacial tension and viscosity was determined experimentally and correlated with the sorption of the gas. This Chapter provides the basis of the process for a better understanding of the different steps during the recycling of PS wastes. 59 Binary Systems 60 Chapter 4 4.1. General Background 4.1.1. Polystyrene/Terpene oils According to the literature, aromatic solvents are efficient for the solubilisation of Polystyrene [1-4], but they are not environmentally friendly and would make difficult the further application of the recycled polymer. The development of a recycling technology must consider the global benefit, sense and coherence of the process; by this reason the screening of the solvents should be carried out according to the principles of the green chemistry. One of the 12 principles of Green Chemistry is concerned about the “use of safer solvents and auxiliaries”. The devastating effect of traditional solvents makes not surprisingly the promotion of alternative solvents [5]. These have low toxicity, are easy to recycle, inert and do not contaminate the final product. Nevertheless, there is no perfect green solvent applicable to all processes. But in this work, we decided to replace the aromatic solvents and/or lubricating oils by terpene oils. The viability of terpenes to solubilise PS was shown in the literature [611], which turns the dissolution of plastic wastes in an environmentally friendly technology. Terpenes are natural compounds originated mainly from plants and in its molecule at least a part is similar to the isoprene subunit, by this reason they are also denoted as isoprenoids. They are present in numerous plants and flowers that smell pleasantly, taste spicy or exhibit pharmacological activities. By these reasons, they have been used throughout history for a broad variety of purposes including perfume, medicine, and flavoring [10]. There are more than 30000 terpenes described in literature, but all of them exhibit the general structure shown below (Figure 4.1). H2C CH2 CH3 Figure 4.1. 2-Methyl-1,3-Butadiene or Isoprene. In nature, terpenes are found mainly as hydrocarbons, alcohols, aldehydes or ketones [12]. In accordance with the wide variety of structures presented by terpenes, the most suitable solvents to carry out the dissolution of PS should present similar structures to the traditional aromatic solvents. By this reason, the monoterpenes were selected as the most adequate. Monoterpenes contain between 10 and 16 atoms of Carbon and are derived from 2,6dimethyloctane. Some of the most well known monoterpenes are /-myrcene, geraniol, o-/m-/p-cymene, linalool, /-phellandrene, terpinolene, -/-terpynene, 61 Binary Systems limonene, cymene, etc. They are present in bosil, bay, hops, roses, lemons, oranges, lavender, mandarins, eucalyptus or orchids, among others. Therefore, it implies that can be obtained from renewable sources. The feasibility of the recycling process could be achieved if the terpenes solvent solubilise selectively the PS, while the rest of the polymers in the waste stream remain insoluble. In general, the rule “like dissolves like” allows the qualitatively comprehension of solubility and miscibility. In general, amorphous polymers can be dissolved in non-polar solvents by mean of ionic and Van der Waals interactions between them. From a thermodynamic point of view, the formation of solutions from pure substances is always favoured by entropy since the spontaneous dissolution process is governed by the free energy of Gibbs [13]. Thus, when two substances are mixed, the disorder of the binary system increases and the enthalpy should be minimized in order to achieve an instantaneously dissolution. The energy of mixing could be represented by the following expression: [4.1] where the left term represents the increment in the cohesive energy per unit of volume; i represents the volume fraction of the substances; 0 the solubility parameter of the solvent and 1’ is the solubility parameter of the polymer (similar to the solubility parameter of the monomer which makes it up) [14]. Thus, solubility parameters () represent the total cohesive energy densities which are the result of several contributions: non polar, dipole-dipole and hydrogen bonds. They represent the numerical values for the application of the rule “like dissolves like” [15]. Thus, Polystyrene will be soluble in those terpenes whose solubility parameters present similar values to itself. But a first screening could be performed according to the chemical structure of the polymer and the solvents. In the case of PS, solvents which exhibit similar solubility parameters will present low tendency to form hydrogen bonds and relatively low polarity. The calculation of could be carried out by approximated [16] or groups contribution methods [13] although most of the solubility parameters of typical solvents are summarized in the literature [15]. The estimation of the solubility parameters of PS and terpenes and the confirmation of its use as predictive tool for the selection of the most suitable solvents is shown below. Also, the effect of temperature, molecular weight and source of the PS wastes was studied experimentally and the results are shown in this section. 62 Chapter 4 4.1.2. Terpene oils/CO2 Compressed supercritical fluids (SCFs) have been proposed as an alternative for several extraction and purification processes since the separation of essential oils can be performed at low temperature which makes them suitable to preserve the thermo labile compounds of the natural oils [17, 18]. The term supercritical fluid is defined as any fluid which is above both its critical temperature and pressure. A fluid in this pseudostate combines properties of gases and liquids. Thus, it can present liquid-like density and no surface tension while interacting with solids surfaces; but also, it has gas-like low viscosity and high diffusivity. Furthermore, its solvent power can be easily tuneable by shifting temperature and pressure [19]. Table 4.1 shows the order of typical properties exhibited by gases, liquids, and supercritical fluids. Table 4.1. Comparison of the physical properties of gases, liquids and supercritical fluids [20]. Gases Supercritical Fluids Liquids Density (kg/m3) 1 100-1000 1000 Viscosity (Pa·s) 10 50-100 500-1000 Diffusivity (mm2/s) 1-10 0.01-0.1 0.001 Supercritical carbon dioxide (scCO2) is by far the most widely SCF use and it is presented as a very interesting substance because its low critical constants (Tc: 31.1 ºC, Pc:73.8 bar) are easily achieved, it is non-reactive, non-toxic, non-flammable, and available at low cost [21]. Although scCO2 has been shown as a promising and suitable technique for the fractionation of natural oils due to its excellent properties, there is an important lack of information concerning the vapour-liquid equilibrium (VLE) data. The solubility dependence of the terpene oils on pressure and temperature forms the basis for the design of the conditions in the separator unit [22], but the complexity of thermodynamic models make difficult the obtaining of generic equilibrium predictions for multicomponent mixtures [23-25]. Because the predictive capabilities of solubility models are rather limited, several authors concluded that it is mostly necessary to determine the solubility experimentally [26]. Although VLE should be obtained experimentally, general trends are observed. Thus, all pure terpenes behaved similarly on all pressures when subjected to the different temperatures [24, 27-30]. By this reason, the solubility of terpenoids in CO2 could be fitted as a function of pressure and temperature in order to generalize and predict the behaviour of essential oils in a wide range of conditions. The aim of this section is to provide a simple semi-empirical model to predict the solubility of essential oils at high pressure on basis of the physico-chemical properties of their compounds. Vapour-Liquid Equilibrium data of binary systems 63 Binary Systems (terpenoids/CO2) from the literature or determined in our research group were used to correlate the semi-empirical model proposed in this work. 4.1.3. Polystyrene/CO2 CO2 has been used in many different applications, such as extraction (as it was shown in the previous section) but mainly in the field of polymers [31-35] since it is non-toxic, non-flammable, inexpensive, presents a solvent strength adjustable and moderate critical temperature and pressure, which makes it an excellent fluid for polymer processing [32]. But the behaviour of amorphous polymers with gas atmospheres is still nowadays a source of scientific interest and a search for new industrial applications [36]. Solubility, diffusivity, density and permeability data are very important for understanding the systems containing a polymer and supercritical CO2 (scCO2) [34]. The gas is diffused between the polymer chains more easily than a large molecule and its sorption increases the free volume and the polymer segment mobility. The absorbed CO2 acts as “lubricant”, making easier for chain molecules to slip over one another and causing polymer softening; by this reason CO2 is known as plasticizer [37]. There are several factors determining the interaction between the plasticizer and the polymers: the chemical nature of the polymer and its physical state, its crystallinity and degree of crosslinking, the intermolecular forces between the polymer and the gas and the molecular size of the diluent (CO 2) [35, 38]. An increase of CO2 concentration promotes swelling and changes in the mechanical and thermal properties of the polymer as a consequence of the swelling and plasticization [36, 39]. As it was mentioned, plasticization causes shifts in the properties of the polymer that results in the decrease of the rigidity at room temperature, depression of the glass transition temperature (Tg), increase of elongation, flexibility and toughness [40, 41]. According to the explained phenomena, for the rational design of processes and applications, the study of the behaviour of the system polymer/CO 2 and the knowledge of the mechanical and physical property changes is crucial. For instance, nucleation of foams using CO2 takes place at temperature close to the Tg [42]. The selection of the most suitable conditions to perform the purification of polymeric matrix is determined by the influence of pressure and temperature on the behaviour of the binary system [37]. The control over the properties and design of functional materials is influenced by the gas sorption into the polymer [36, 43]. Also, reliable high pressure rheological data are essential for the application of CO 2 to induce plasticization in the industrial plastics processes [40, 44]. The gas solubility in the polymer can be described by Henry’s law: c=H·P 64 [4.1] Chapter 4 where c is the concentration of the gas in the polymer, H is Henry’s law constant and P is the equilibrium gas pressure. Henry’s law is only valid on the ideal solution state and in the case of diluted solutions because it does not consider the interactions between the gas and the polymer. Although, it is not recommended its use at high pressure, it has been widely used due to its simplicity. As an alternative, CO2 sorption in amorphous polymers is described using the Dual-Mode model [36, 45]. Dual Mode Sorption Model is composed of Henry’s law dissolution in the equilibrium region (low pressures) and Langmuir-type sorption in a non-equilibrium region [46]. The non-equilibrium region is related to the excess free volume or unrelaxed free volume in a glassy polymer resulting from the presence of microcavities capable of retaining solute molecules[47]. The Dual Mode sorption model describes the solubility in the isotherm according to the expression shown below: [4.2] where S is the sorption of CO2 in the polymer, kH is analogous to Henry’s law constant, P is the pressure, C’H is the saturation of the cavities and b represents the affinity between the solute molecules and the Langmuir sites present in the polymeric matrix. On the other hand, the Sánchez-Lacombe Equation of State (EoS) has been also used to correlate the sorption of CO 2 in PS [48] since it is the most widely used thermodynamic model to describe the behaviour of polymer/gas solutions. It is based on the classical lattice-fluid theory extended with vacancies in the lattice in order to account for compressibility. The equation of state for pure fluids is shown below: [4.3] where , , and represents the reduced density, pressure and temperature, respectively. The reduced parameters are defined from characteristic fluid parameters: [4.4] [4.5] [4.6] [4.7] [4.8] [4.9] [4.10] [4.11] [4.12] [4.13] where P*, T*,* and v* are the characteristic pressure, temperature, close-packed mass density and molar volume, respectively and they are typically fit to vapour pressure or liquid density data; is the closest distance allowed between two mers, is the mer interaction energy, v0 is the hole volume, R is the gas ideal constant, MW is the molecular weight, v* is the hard-core molecular volume and r is a size parameter. The lattice fluid model for mixtures is obtained from the application of the mixing rules. 65 Binary Systems [4.14] [4.15] [4.16] where i is the close-packed volume fraction of component i in the mixture. It is related to mass fraction (wi) by: [4.17] The parameter Xij is calculated from a geometric mean: [4.18] where kij is an adjustable binary interaction parameter essential in accurately modelling the mixture behaviour, and on understanding the affinity of gas molecules to the polymer. It is obtained from the correlation of experimental data and its aim is the correction for the deviation of the mixture. Results have been divided in two main blocks. Initially, the sorption of CO 2 was determined experimentally and correlated using Sánchez-Lacombe and Dual-Mode model. Next, the influence of the CO 2 sorption on the modification of the PS properties was studied. The sorption of CO 2 in Polystyrene produces polymer swelling [49-51], large depression in glass transition temperatures [42, 52, 53], interfacial tension [54, 55] and viscosity [56]. 4.2. Polystyrene/Terpene oils Results 4.2.2. Determination of the solubility parameters Many methods have been developed for the determination of the solubility parameter, from theoretical calculation, to empirical correlations [16]. In this work, the groups contribution method was selected because it allows the calculation of the partial contribution to the solubility parameters, which helps to a better understanding than the global solubility parameter. Thus, the global solubility parameter is composed by: [4.19] where d is the energy from dispersion forces, p is the energy from dipolar intermolecular forces and h is the energy due to the hydrogen bonds between molecules. There are several authors who estimated the different groups contribution to the solubility parameters. Among the best well known, are Small [57], Hoy, van Krevelen and van Krevelen and Hoftyzer [58, 59]. In this work, 66 Chapter 4 values shown by Hoy were selected to calculate the individual solubility parameters because they fitted pretty well the estimation and the experimental shown in the literature for a few terpenes. Furthermore, it allows the calculation of the partial solubility parameters of the Polystyrene since his method included a corrective term for non-ideality, to be applied in the case of polymers [60]. Table 4.2 shows the predicted solubility parameters for the selected terpene oils according to Hoy’s method. Table 4.2. Solubility partial parameters of the selected terpene oils. Solvent d(MPa1/2) p(MPa1/2) h(MPa1/2) Anisole 16.82±0.98 10.15±6.05 9.14±2.44 p-Cymene 17.95 6.88 4.59 Cinnamaldehyde 15.80±3.60 12.10±0.30 10.26±4.06 Eucalyptol 15.01±0.89 7.41±3.51 9.08±5.68 -Phellandrene 15.93 5.68 7.14 Geraniol 14.76 8.38 12.11 Limonene 15.93 4.85 6.59 Linalool 14.02±2.28 7.72±3.32 11.28±0.08 -Pinene 16.48 6.87 7.33 -Terpinene 16.45 5.88 6 14.08±0.08 7.47±0.53 11.72±1.52 -Terpineol Experimental values 1 -Terpineol 2 Cinnamaldehyde 3 Eucalyptol 4 Linalool d(MPa1/2) 14.00 19.40 15.90 16.30 p(MPa1/2) 8.00 12.40 3.90 4.40 h(MPa1/2) 10.20 6.20 3.40 11.20 67 Binary Systems According to Table 4.2 the predicted and experimental vales of d and h are very similar; however the most important deviations are due to the energy due to the hydrogen bonds. In order to establish a relationship between the solubility parameters of the terpene and the solubility of PS, Hansen proposed that a polymer will be soluble in a solvent if of the solvent is within its solubility sphere [15]. Thus, the knowledge of the radius and the solubility parameters of the Polystyrene are necessary in order to fix the central point of the sphere. The establishment of the solubility parameter of PS is not so evident since a wide variety of characteristics are involved in its structure (molecular weight, polidispersity index, crystallinity, tacticity or crosslinking) and consequently, they will influence on its solubility. According to the literature, the solubility parameters shown by Szydlowski et al. [61] and the interaction ratio calculated by Bernardo and Veseley [62] presented similarities to the PS used in this work (Table 4.3). Table 4.3. Solubility partial parameters of Polystyrene from the literature [61]. Polystyrene d(MPa1/2) p(MPa1/2) h(MPa1/2) 17.90 4.20 5.00 Figure 4.2 shows the graphical methodology established by Hansen [15] to predict the ability of terpenes to dissolve PS, where the markers represent the intersection of the partial solubility parameters and the sphere the interaction ratio of the polymer. p-Cymene 15 -Terpineol Menthol Cinnamaldehyde -Pinene -Phellandrene Eucalyptol Geraniol Limonene Linalool -Terpinene Anisole Polystyrene 12 p 9 6 3 0 0 5 10 15 d 20 25 3 30 0 6 9 12 15 h Figure 4.2. Prediction of PS solubility in terpene oils from the solubility parameters shown in Table 4.2 for terpenes. As it was mentioned, the ability of solvents to dissolve a polymer is indicated by the proximity between the polymer and the studied solvent. In Figure 4.2, the solubility parameter of PS is placed in the centre of the sphere of radius h. If the solubility 68 Chapter 4 parameters of the terpene solvents remain inside the sphere, the terpene is considered as good solvent. Whereas, if the solubility parameters of the solvents are placed outside, it is may be assumed that the studied terpene will not dissolve the polymer [63]. According to Figure 4.2, the most suitable solvents to perform the recycling of PS by dissolution are p-Cymene, Limonene, -Terpinene, Phellandrene and -Pinene. The solubility of PS in the selected monoterpenes was determined experimentally to check the feasibility of the theoretical methods for the prediction of the polymer dissolution. Table 4.4. shows the values of experimental solubility of PS in the set of terpene oils at 25 ºC. Table 4.4. Experimental solubility of PS in terpene oils at 25 ºC. Solvent Solubility (g/ml) Solvent Solubility (g/ml) p-Cymene 0.3529 -Terpineol 0.0063 Anisole 0.2578 -Pinene 0.2901 Cinnamaldehyde 0.1957 -Phellandrene 0.2879 Linalool 0.0059 Geraniol 0.0029 -Terpinene 0.2427 D-Limonene 0.2473 Eucalyptol 0.2977 According to Table 4.4, the most suitable solvent to carry out the dissolution process agrees with the predicted theoretically (Figure 4.2). p-Cymene is the most suitable solvent to carry out the dissolution process, because presents the greatest solubility values. Moreover to dissolve easily the PS, a good solvent to carry out its recycling should show high volatility, which allows its removal from the polymer. Also it is highly appreciated that the terpenes present low cost, low toxicity and easy availability [64]. The strong interactions between PS and some of the cited solvents (those who exhibited high solubility values) could prevent its complete removal from the polymer-rich solution [65, 66]. According to the described reasons, the most suitable terpene oils for the recycling of PS wastes should be: p-Cymene, Terpinene, Limonene, -Pinene and -Phellandrene. The subsequent sections will be focused on the dissolution of PS in the mentioned terpene oils. 4.2.3. Influence of the Molecular Weight on the Solubility of Polystyrene The molecular weight of a polymer is an intrinsic property with a great influence on the properties of their solutions. The activity coefficient at infinite dilution (Ω ∞1) could be useful to predict the phase equilibrium over the whole concentration range of a polymer solution. The influence of molecular weight on the activity coefficient at infinite dilution is predicted using Equation 4.20 [67]. 69 Binary Systems ln 1 1 MW1 MW1·12 MW2 [4.20] where MW1 is the molecular weight of the solvent, MW2 is the molecular weight of the polymer and 12 is the binary interaction parameter, which shows the physical interaction between polymer and solvent, it was optimized from experimental solubility data, obtaining a value of 0.3220. If experimental data are not available, 12 could be approximated to 1 [15], particularly for systems where dispersion forces dominate over polar and hydrogen-bonding ones, but it systematically overestimates the infinite dilution solvent activity coefficients [68]. This indicates that better results may be obtained with lower 12 values as we present. From a theoretically point of view, the relationship between PS molecular weight and the activity coefficient at infinite dilution should be approximately constant in the range of molecular weight studied. Figure 4.3 shows the theoretical relationship between activity coefficients at infinite dilution versus MW, the interval includes from styrene monomer (104.1 Da) to 108 Da molecular weight. 4.0 3.5 3.0 2.5 2.0 p-Cymene -Terpinene D-Limonene -Pinene -Phellandrene 1.5 1.0 0.5 10 1 10 2 10 3 10 4 10 5 10 6 10 7 10 8 10 9 Molecular Weight (g/mol) Figure 4.3. Evolution of the activity coefficient at infinite dilution as a function of Molecular Weight. Figure 4.3 can be divided in two regions. the first one belongs to low molecular weight (100-5000 g/mol) where the increase of Ω∞1 is directly proportional to the polymer molecular weight [69, 70], and the second corresponding to higher molecular weights (> 10000 g/mol), where Ω∞1 remains practically constant. According to the observed trend, solubility does not decrease when polymer molecular weight increases in the range of study [71, 72]. 70 Chapter 4 In order to determine the real effect of PS molecular weight on the solubility in terpenes, three commercial samples were used. Generally, the average molecular weight of the common PS available in plastics is ranged between 150000 and 400000 g/mol [73]; in this work the samples had molecular weight values of 192000, 280000 and 350000 g/mol (Figure 4.4). 0.45 Solubility (g/ml) Mw: 192000 g/mol 0.40 Mw: 280000 g/mol 0.35 Mw: 350000 g/mol 0.30 0.25 0.20 0.15 0.10 0.05 0.00 p-Cymene Terpinene Limonene Pinene Phellandrene Terpene solvents Figure 4.4. Experimental solubility determined as a function of the Molecular Weight of PS at 25 ºC. Figure 4.4 shows the effect of molecular weight on the solubility of PS in the previously selected terpenes. It is observed that solubility remains practically constant over the whole studied range since the theoretical prediction showed that molecular weight does not influence significantly from 105 g/mol. By this reason, several authors have established that molecular weight is not the most relevant variable during the polymer dissolution process. 4.2.4. Influence of the Temperature on the Solubility of Polystyrene It is generally known that an increase on temperature produces an increase in the solubility. This fact is explained according to the thermodynamic rule which says that a substance is soluble in a solvent if the Gibbs free energy is negative. Flory– Huggins theory [74, 75] describes the thermodynamic equilibrium for polymer and solvent binary systems at constant pressure, as it is shown in Equation 4.21. 2 Gm 1 R T V ln 1 V ln 2 121 2 1 2 [4.21] 71 Binary Systems where T is the temperature, R is the ideal gas constant, is the volume fraction of each component in the mixture, V is the molar volume of the two components, 12 is the Flory–Huggins interaction parameter expressing the interaction enthalpy between two different molecules, the subscript 1 represents the solvent and 2 represents the polymer. The first two terms on the right side of this equation denote the entropy of mixing. which quantitatively is negative[76], since in solution the molecules display a more chaotic arrangement than in the solid state. Enthalpy of mixing (ΔHm) at constant pressure for polymer–solvent mixtures can be calculated as: H m 1212 [4.22] where the 12 parameter expresses the interaction enthalpy between two different molecules. This parameter is always positive, opposing the negative entropy effect of mixing [77], thus the lower values of 12 show greater results in the solubility as it can be checked in Figure 4.5, where the influence of temperature over 12 is depicted. 0.8 0.14 0.12 p-Cymene -Terpinene D-Limonene -Pinene -Phellandrene 0.7 0.6 Solubility (g/ml) 0.10 X12 p-Cymene -Terpinene D-Limonene -Pinene -Phellandrene 0.08 0.06 0.04 0.5 0.4 0.3 0.2 0.02 0.1 0.00 0.0020 0.0025 0.0030 1/T (K-1) 0.0035 20 30 40 50 60 70 Temperature (ºC) Figure 4.5. (Left) Prediction of the influence of the temperature on 12. (Right) Effect of temperature on PS pellets solubility obtained experimentally. As it is observed in Figure 4.5 (left). there is a linear relationship between 12 and the inverse of temperature [78]. Accordingly, when the temperature increases, the value of 12 decreases, and therefore, as the temperature increase, dissolution process is favoured. 72 Chapter 4 Furthermore, the effect of temperature on the solubility was investigated experimentally in the temperature range from 25 to 60 ºC (Figure 4.5 right), higher temperatures were not used to avoid polymer degradation. PS sample selected was 280000 g/mol because as it was explained it is not the most important characteristic to carry out the dissolution process, so it was decided to select the average molecular weight among the previously tested samples. As it is observed, an increase on temperature produces an increase in the solubility in all the solvents studied. Nevertheless, the enhancement of solubility due to the temperature increase should be compared with the energetic costs in order to decide the most economically valuable way to dissolve the PS. In this work, room temperature or slightly higher were selected as the most convenient to manage PS wastes by dissolution. 4.2.5. Applicability to real Polystyrene wastes Polystyrene is processed on different ways to adapt the market needs. Generally, PS wastes can be found as expanded (EPS) or extruded (XPS) Polystyrene, which provide a wide applicability field. The processing of polymers could affect to their dissolution in terpene oils since the dissolution process can be affected by the chain chemistry, composition and stereochemistry [50]. The dissolution of a polymer into a solvent is not immediate and involves two transport processes, namely solvent diffusion and chain disentanglement. Initially, the solvent begins its aggression by pushing the swollen polymer substance into the solvent, and, as time progresses, a more dilute upper layer is pushed in the direction of the solvent stream. Further penetration of the solvent into the solid polymer increases the swollen surface layer until, at the end of the swelling time, a quasistationary state is reached where the transport of the macromolecules from the surface into the solution prevents a further increase of the layer [51]. The way in which the wastes are processed could affect the dissolution process. Since the previous parts of this study were performed using pellets of PS, in this section, the influence of different sources on the solubility of PS in terpene oils was evaluated (Figure 4.6.) 73 Binary Systems 0.40 XPS PS Pellets EPS 0.35 Solubility (g/ml) 0.30 0.25 0.20 0.15 0.10 0.05 0.00 p-Cymene Terpinene Limonene Pinene Phellandrene Terpene solvents Figure 4.6. Influence of the origin on the solubility of PS and PS wastes in terpene oils at 25 ºC. According to Figure 4.6, the solubility of PS from different sources remains almost constant independently of their origin. As it was observed in the previous section entitled “Influence of Molecular Weight”, when the molecular weight of the polymer exceeds 105 g/mol, does not affect on the solubility. In the case of commercial polymers, particularly PS, molecular weight is always higher than the mentioned value and consequently, solubility remains almost constant, as it was expected. Nevertheless, differences regarding the dissolution process were observed and by this reason, the minimum time required to dissolve PS pellets was studied. Figure 4.7 shows the thermograms which represent the amount of polymer solubilised in Limonene in the samples withdrawn at different times from 10 seconds to 24 hours. According to Figure 4.7, the amount of polymer dissolved in the case of PS (left) increases slightly during the first 60 minutes, but 24 hours are necessary to solubilised fully the pellets. Similar behaviours were observed with the others terpene oils, which required almost 24 hours to dissolve completely PS pellets due to its high compression grade. On the other hand, Figure 4.7 right represents the amount of XPS dissolved in Limonene. 74 Chapter 4 100 80 70 60 XPS 10 sec XPS 30 sec XPS 60 sec XPS 5 min XPS 15 min XPS 60 min XPS 24 h 90 80 70 Weight (%) 90 Weight (%) 100 PS 10 sec PS 30 sec PS 60 sec PS 5 min PS 15 min PS 60 min PS 24 h 50 40 60 50 40 30 30 20 20 10 10 0 0 0 100 200 300 400 500 0 100 Temp (ºC) 200 300 400 500 Temp (ºC) Figure 4.7. Evolution of PS and XPS solubilised in Limonene by thermogravimetric analysis. Temperature: 25 ºC; stirring: 500 rpm. The dissolution of XPS wastes increases gradually when time increases, although 24 hours are required to achieve a fully solubilisation. Billmeyer [14] explained that the dissolution process of the polymers as a slow process that occurs in two stages. Figure 4.8 summarizes the evolution of the solubility of PS pellets as a function of time in the five terpene oils. The tendency shows that during the initial 100 seconds PS in pellets is practically insoluble in the solvent, but since that time, the dissolution process starts following an exponential behaviour. 0.5 0.5 PS pellets p-Cymene -Terpinene D-Limonene -Pinene -Phellandrene 0.3 0.4 Solubility (g/ml) Solubility (g/ml) 0.4 XPS wastes p-Cymene -Terpinene D-Limonene -Pinene -Phellandrene 0.2 0.1 0.3 0.2 0.1 0.0 1 10 100 1000 Time (s) 10000 100000 1 10 100 1000 10000 100000 Time (s) Figure 4.8. (Left) Time influence in PS pellets solubility and XPS wastes solubility (right). The higher porosity and surface area of the foam make easier the sorption of the solvent in the polymer, and hence the dissolution process is faster. From the gathering of Figures 4.6, 4.7 and 4.8 is possible to assert that results obtained for PS 75 Binary Systems pellets are applicable to XPS and EPS. Firstly, solvent molecules slowly diffuse into the polymer to produce a swollen gel. This may be all that happens if, for example, the polymer-polymer intermolecular forces are high because of crosslinking, crystallinity, or strong hydrogen bonding. But if these forces can be overcome by the introduction of strong polymer-solvent interactions and the second stage of solution can take place. Here the gel gradually disintegrates into a true solution and only this stage can be materially speeded by agitation [79]. To sum up, terpene solvents present excellent characteristics for PS recycling due to their low toxicity, cost and high solubility, that has been checked experimental and theoretically. Moreover, in the range of molecular weight studied, the solubility of the polymer remains constant independently of it and as it was expected, temperature favours the solubility of the polymer. Finally, theoretical results can be applied to real wastes which confirm the feasibility of the dissolution of PS wastes in terpenes as initial step for its recycling. 4.3. Terpene oils/CO2 Results 4.3.1. Procedure The vapour phase was correlated following the traditional Chrastil’s equation [80] which is based on the hypothesis that each molecule of solute (terpene) associates with k molecules of supercritical solvent (CO2) to form a solvate-complex, which is in equilibrium with the system. [4.23] where S is the solubility in [kg/m3], C1 is a constant dependent on the molecular weights of the solute and solvent and on the association constant, C2 is a constant dependent on the total heat of vaporization, T is the temperature in [K], k is the association number of molecules in the solvate-complex and is the CO2 density in [kg/m3]. The new empirical relationship proposed in this work for the correlation of the liquid phase was: [4.24] where xCO2 is the mole fraction of CO2 in the liquid phase, Ai are the correlated variables, P is the pressure in [bar] and T is the temperature in [K]. Initially, the calibration of Ai, Ci and k was carried out from experimental data obtained from the literature or in our research group (Table 4.4). They were calculated by performing a multiple linear regression and minimising the sum of the square differences between experimental and calculated solubility. Although there are some binary systems well described in the literature, there is also an important 76 Chapter 4 deficiency on VLE data of minority or non common compounds of the essential oils. The presence of the cited components could modify significantly the solubility of the terpenoids in CO2. By this reason, the fitted constants (Ai, Ci and k) were related with the physical properties of the terpenes in order to get an empirical relationship which allows the prediction of new components whose VLE data in CO2 were not previously described in the literature. Table 4.5 shows a good sample of the available data from the literature about the equilibrium of the selected terpenoids in CO2 as a function of pressure and temperature. VLE measurements of terpenoids were carried out in a pressure range from 7.8 to 133.4 bar and in a temperature range from 303 to 343 K. Different methods could be employed to determine the vapour-liquid equilibrium for the binary systems containing CO2/terpenoids, but according to the literature the static method was the most commonly used [24, 27-29]. Table 4.5. Chemical structure, range of pressure, temperature, experimental method and source for the binary system terpenoids/scCO 2. Temperature (K) range Pressure Number (bar) range of Data Carvone Method Ref 303.55-324.05 18-96 30 Dynamic VLE Cell with stirring [27] 303.15-343.15 1.9-110.7 27 Static VLE Cell with stirring This work 318.2-323.2 77.5-98.0 15 Static VLE Cell [28] 313.18-333 7.8-110.6 29 Static VLE Cell with recirculation [29] 313.78-333.25 8.3-110.0 29 Static VLE Cell with recirculation [29] 313-333 62.9-109.3 24 Static VLE Cell [24] p- Cymene Eucalyptol Linalool Limonene -Pinene 77 Binary Systems Table 4.5 (Continued). Chemical structure, range of pressure, temperature, experimental method and source for the binary system terpenoids/scCO 2. Temperature (K) range Pressure Number (bar) range of Data -Terpinene 313.15-343.15 4.9-110.5 26 Method Ref Static VLE Cell with stirring This work This semi-empirical relationship, proposed in this work, provides an important tool for the generalization and prediction of minority species which will enhance the estimation of the global phase behaviour of the essential oils. The equation which allows the calculation of Ai, Ci or k constants from the physical properties of the terpenoids could be generalized for the three constants: [4.25] where Zi means Ai, Ci or k; zi,j means ai,j, ci,j or ki (respectively) and the subscript i means 1 or 2 and the subscript j varies from 1 to 5. The properties of the terpenes are: MW is the molecular weight in [g/mol], Tb is the boiling temperature in [K], is the density in [kg/m3] and is the solubility parameter calculated following the method previously described in the literature and in the previous section of this Chapter [13]. Once the composition of the essential oils is defined, the prediction of the vapourliquid equilibrium data for the essential oils in carbon dioxide could be achieved with the simply knowledge of the described properties. The essential oil solubility will be the weighted average of the individual solubility of each of the terpene in CO2. The global procedure is described in Figure 4.9. 78 Chapter 4 Figure 4.9. Process scheme to correlate and predict essential oil equilibrium data. Validity of the predictions was checked by comparison with literature data concerning ternary systems (Limonene + Linalool + CO 2) and real mixtures (Lemon and orange oils + CO2) which contain compounds whose VLE data are not described in the literature. 4.3.2. Literature Terpenoids/CO2 data and correlation of binary systems Essential oils are a complex mixture of terpenes or their derived compounds but data for natural and well characterized oils mixtures with supercritical CO 2 are scarce in the literature. Our approach considered that essential oils consisted mainly of Limonene, Linalool, -Terpinene, Carvone, -pinene, p-Cymene and Eucalyptol. It is necessary to emphasise that large database for CO 2 and the cited terpenoids are available in the literature which could enhance the accuracy of the calibration of the constants, but also the dispersion could be responsible of deviations in the estimation of parameters Figures 4.10-4.16 show the experimental data (markers) and the correlated values (lines) of solubility of binary systems CO2/terpenoids at different temperatures. From the correlation of the experimental data following eq. [4.6] and [4.7], the 79 Binary Systems constants Ai, Ci and k were obtained. Table 4.6 shows the values for the fitting constants employed in the solubility modelling of terpenoids using the semiempirical equation. In all cases, the deviations were below the constants values. 120 20 T: 303.5 K T: 313.5 K T: 323.5 K 16 Solubility (kg Carv/m CO2) 100 T: 313.5 K 18 14 Pressure (bar) 3 80 60 40 20 12 10 8 6 4 2 0 0 0.0 0.2 0.4 0.6 0.8 1.0 0 xCO2 50 100 150 200 250 300 3 Density of CO2 (kg/m ) Figure 4.10. Solubility of Carvone in CO2. Markers represent experimental data at 303.5 K (■), 313.5 K (●) and 323. 5 K (▲) obtained from [27] and lines the correlation at 303.5 K (—), 313.5 K (—) and 323.5 K (—). 120 20 T: 313.15 K T: 323.15 K T: 343.15 K Solubility (kg p-Cym/m CO2) 100 16 14 3 80 Pressure (bar) T: 313.15 K T: 323.15 K T: 343.15 K 18 60 40 20 12 10 8 6 4 2 0 0 0.0 0.2 0.4 0.6 xCO2 0.8 1.0 0 50 100 150 200 250 300 350 400 450 3 Density of CO2 (kg/m ) Figure 4.11. Solubility of p-Cymene in CO2. Markers represent experimental data at 313.5 K (■), 323.5 K (●) and 343. 5 K (▲) and lines the correlation at 313.5 K (—), 323.5 K (—) and 343.5 K (—). 80 Chapter 4 120 35 T: 318.2 K T: 323.2 K 30 Solubility (kg Euc/m CO2) 110 100 25 3 Pressure (bar) T: 318.5 K T: 323.5 K 90 80 70 15 10 5 0 20 15 10 5 0 0.0 0.2 0.4 0.6 0.8 1.0 0 xCO2 50 100 150 200 250 300 350 400 3 Density of CO2 (kg/m ) Figure 4.12. Solubility of Eucalytpol in CO2. Markers represent experimental data at 318.2 K (■), and 323.2 K (●) from [28] and lines the correlation at 318.2 K (—), 323.5 K (—). 20 T: 313.8 K T: 323.2 K T: 333.2 K 120 16 Solubility (kg Lim/m CO2) 100 3 80 Pressure (bar) T: 313.8 K T: 323.2 K T: 333.2 K 18 60 40 20 14 12 10 8 6 4 2 0 0 0.0 0.2 0.4 0.6 xCO2 0.8 1.0 0 50 100 150 200 250 300 350 400 450 3 Density of CO2 (kg/m ) Figure 4.13. Solubility of Limonene in CO2. Markers represent experimental data at 313.8 K (■), 323.2 K (●) and 333.2 K (▲) and lines the correlation at 313.8 K (—), 323.2 K (—) and 333.2 K (—). 81 Binary Systems 20 T: 313.8 K T: 323.2 K T: 333.2 K 120 16 Solubility (kg Lin/m CO2) 100 3 80 Pressure (bar) T: 313.8 K T: 323.2 K T: 333.2 K 18 60 40 14 12 10 8 6 4 20 2 0 0 0.0 0.2 0.4 0.6 0.8 1.0 0 xCO2 50 100 150 200 250 300 350 400 450 3 Density of CO2 (kg/m ) Figure 4.14. Solubility of Linalool in CO2. Markers represent experimental data at 313.8 K (■), 323.2 K (●) and 333.2 K (▲) from [29] and lines the correlation at 313.8 K (—), 323.2 K (—) and 333.2 K (—). 120 20 T: 313 K T: 323 K T: 333 K T: 313 K T: 323 K T: 333 K 18 16 14 Pressure (bar) 3 Solubility (kg -Pin/m CO2) 100 80 60 15 10 5 0 12 10 8 6 4 2 0 0.0 0.2 0.4 0.6 xCO2 0.8 1.0 0 50 100 150 200 250 300 350 400 450 3 Density of CO2 (kg/m ) Figure 4.15. Solubility of -Pinene in CO2. Markers represent experimental data at 313 K (■), 323 K (●) and 333 K (▲) obtained from [24] and lines the correlation at 313 K (—), 323 K (—) and 333 K (—). 82 Chapter 4 120 20 T: 313.15 K T: 323.15 K T: 343.15 K Solubility (kg -Terp/m CO2) 100 16 14 3 80 Pressure (bar) T: 313.15 K T: 323.15 K T: 343.15 K 18 60 40 20 12 10 8 6 4 2 0 0 0.0 0.2 0.4 0.6 0.8 1.0 0 50 100 150 200 250 300 350 400 450 3 xCO2 Density of CO2 (kg/m ) Figure 4.16. Solubility of -Terpinene in CO2. Markers represent experimental data at 313.15 K (■), 323.15 K (●) and 333 K (▲) and lines the correlation at 313.5 K (—), 323.5 K () and 343.15 K (…). According to Figures 4.10-4.16, it is observed that the semi-empirical model tested for the liquid and vapour phase fit well the experimental results, but the model of Chrastil for the vapour phase presented less standard deviation than the semiempirical equation which describes the liquid phase. Table 4.6. Fitted constants and associated error from the correlation of the semi-empirical model for the selected terpenoids. Terpenoid Carvone p-Cymene Eucalyptol Linalool Limonene -Pinene -Terpinene A1 43.33 ± 1.46 38.51 ± 1.03 32.84 ± 0.71 29.07 ± 0.60 29.80 ± 0.80 29.80 ± 1.02 37.53 ± 1.02 A2 -2.23·10-5 ± 8.87·10-6 -1.41·10-5 ± 3.52·10-6 -2.94·10-6 ± 2.03·10-6 -2.18·10-6 ± 6.88·10-7 -2.38·10-6 ± 1.08·10-6 -2.97·10-6 ± 1.50·10-6 -2.97·10-6 ± 6.14·10-7 83 Binary Systems Table 4.6 (Continued). Fitted constants and associated error from the correlation of the semi-empirical model for the selected terpenoids. Terpenoid C1 C2 k Carvone p-Cymene Eucalyptol Linalool Limonene -Pinene -Terpinene -32.33 -4.82 ± 6.22 -34.86 ± 8.80 -22.01 ± 2.73 -6.68 ± 1.65 -14.63 ± 4.38 -4.82 ± 6.22 2549.93 -3462.09 ± 1534.80 -10442.96 ± 3765.92 3024.05 ± 742.38 -1902.82 ± 456.42 -3127.34 ± 860.00 -3462.09 ± 1534.88 4.78 3.06 ± 0.62 11.97 ± 1.71 2.56 ± 0.21 2.57 ± 0.20 4.58 ± 0.40 3.06 ± 0.62 The values of Ai, Ci and k constants were in an enclosed range since they belong to the same compounds family which implies similar structures and properties. Nevertheless, some discrepancies were observed among the constants values of the different terpenoids, which suggested the classification of data into two groups. Attending to the chemical structures it could be observed that the presence of an oxygen atom entails higher molecular weight and boiling point which leads lower solubility of the terpene in CO2 due to the increase in polarity. Following this argument, we propose the classification of terpenoids in oxygenated (Linalool, Carvone and Eucalyptol) and non oxygenated compounds (Limonene, -Terpinene, -Pinene and p-Cymene) in order to get the fitting parameter which correlate the physico-chemical properties of the compounds with Ai, Ci or k. The physico-chemical properties required for the correlation of eq. [4.25] are shown in Table 4.7. Table 4.7. Physico-chemical properties of the terpenoids. MW: molecular weight; T b: boiling temperature; density; solubility parameter. MW (g/mol) Carvone p-Cymene Eucalyptol Linalool Limonene -Pinene -Terpinene 150.22 134.22 154.25 154.25 136.24 136.23 152.23 Tb (K) (kg/m3) (MPa)1/2 504.15 450.15 449.15 471.65 449.65 428.65 458.15 960.00 857.00 922.50 863.00 841.10 858.00 850.00 19.70 19.76 19.04 19.58 17.91 19.30 18.47 Table 4.8 shows the values obtained from the fitting of the Zi (Ai, Ci and k) constants using eq. [4.25] attending the classification of oxygenated and non oxygenated terpenes. 84 Chapter 4 Table 4.8. Correlation constants obtained from the Equation [4.25] for oxygenated and non oxygenated terpenes. The Average Standard Deviation (ASD) of the correlated and predicted values for each constant was shown. Non Oxygenated Terpenes A1 a1,1 a1,2 a1,3 a1,4 a1,5 ASD 0.275 0.228 -0.285 7.036 2.921 2.342·10-4 C1 c1,1 c1,2 c1,3 c1,4 c1,5 ASD A2 a2,1 a2,2 a2,3 a2,4 a2,5 ASD 1.360 -0.851 -3.630 30.796 2696.852 1.618·10-2 C2 -0.095 0.396 -0.289 3.774 3.614 8.112·10-4 c2,1 c2,2 c2,3 c2,4 c2,5 ASD k -142.010 3.709 49.009 -1421.616 21.551 0.330 k1 k2 k3 k4 k5 ASD 0.088 -0.074 0.023 0.054 3.660 2.767·10-4 Oxygenated Terpenes A1 a1,1 a1,2 a1,3 a1,4 a1,5 ASD -0.645 0.071 0.096 0.632 3.613 2.126·10-3 A2 a2,1 a2,2 a2,3 a2,4 a2,5 ASD 17.469 1.626 -0.747 -144.885 2702.536 8.806·10-3 Oxygenated Terpenes C1 C2 k c1,1 c1,2 c1,3 c1,4 c1,5 0.129 0.150 -0.152 0.767 3.655 c2,1 c2,2 c2,3 c2,4 c2,5 -245.203 296.389 -114.363 -12.908 2.506 k1 k2 k3 k4 k5 0.009 -0.200 0.089 0.751 3.654 ASD 2.779·10-3 ASD 1.505 ASD 1.960·10-3 The fit of the constants as a function of the molecular weight, the boiling temperature, the density and the solubility parameter of the terpenoids is so accurate that the values of solubility obtained overlay the correlation data showed 85 Binary Systems in Figures 4.10-4.16. The low values of the standard deviation show in Table 4.8 confirmed that the fit of Ai, Ci and k following eq. [4.25] is highly accurate. 4.3.3. Prediction of multicomponent mixtures and essential oils The feasibility of the prediction of the VLE data of terpenoids in CO 2 was checked using the mixture of Limonene and Linalool. The ternary system Limonene/Linalool/CO2 has been studied since they are considered the key compounds of the citrus oil, and they are the most difficult compounds to separate [81, 82]. The literature data was obtained from Cháfer et al. [81] and the comparison with the prediction is shown in Table 4.9 at 318.15 and 328.15 K in a pressure range between 70 and 110 bar. Table 4.9. Comparison between experimental and predicted solubility of the ternary system Limonene/Linalool/CO2 at 318.15 K and 328.15 K. P (bar) 70 74 78 82 86 90 70 74 78 82 86 90 95 100 105 110 SLim (kg/m3) SLim (kg/m3) Std. SLin (kg/m3) SLin (kg/m3) Std. Exp. Pred. Dev. Exp. Pred. Dev. Temperature: 318.15 K 2.070 2.163 0.066 0.649 2.401 1.239 2.748 3.012 0.187 1.108 3.339 1.578 3.522 4.039 0.365 1.574 4.473 2.050 8.716 6.047 1.887 6.524 6.687 0.115 25.387 9.630 11.142 17.888 10.630 5.132 83.621 14.279 49.032 57.308 15.738 29.394 Temperature: 328.15 K 4.089 1.896 1.551 2.305 1.315 0.701 4.619 2.432 1.546 2.540 1.685 0.604 5.107 3.053 1.452 2.844 2.113 0.517 6.016 3.882 1.509 3.505 2.685 0.580 7.453 4.976 1.751 4.451 3.438 0.716 10.612 6.240 3.091 5.691 4.308 0.978 19.550 8.900 7.530 7.807 6.135 1.182 32.958 12.158 14.708 18.245 8.371 6.982 54.791 15.294 27.928 33.411 10.521 16.186 133.848 18.882 81.293 63.947 12.978 36.040 Table 4.9 shows that the proposed semi-empirical model predicts pretty well the solubilities of Limonene and Linalool in CO2, although higher deviations were observed at higher pressure. However in the range of pressure between 7 and 9 MPa the prediction was accurate. These pressure are particularly relevant for the natural oils rich in terpenoids since they represent potential working conditions for 86 Chapter 4 the efficient application of the supercritical processes [83, 84]. The predicted values are underestimated, but according to the VLE data used for correlation [29], the maximum solubility of the terpenoids at 323.15 K was 20 kg Limonene/m3 CO2 and 15 kg Linalool/m3 CO2. Experimentally the mixture reached solubility values of 133 kg Limonene/m3 CO2 and 63 kg Linalool/m3 CO2, the increase of solubility in the ternary mixture was attributed by some authors to the interactions between the compounds [30]. With the aim of validating the semi-empirical model, the prediction of the solubility of natural oils in CO2 was carried out [84, 85]. The Lemon and Orange oils contain terpenoids whose VLE data have not been described in the literature and the prediction of Ai, Ci and k constants based on their properties will be very useful to predict their behaviour. In the case of Lemon oil, Limonene and Citral were assumed as the majority compounds and in the Orange oil, the prediction of solubility was based on the Limonene, Linalool and Geranial. The solubility of Lemon oil in CO2 was initially predicted. Gironi and Maschietti determined the composition of the Lemon oil by means of Gas Chromatographic analysis and concluded that the main classes of compounds were: monoterpenes (93.7 wt.%), monoterpene oxygenated derivatives (4.3 wt.%) and sesquiterpenes (2 wt.%) [84]. In this work, we considered that the lemon oil could be modelled using Limonene as monoterpene and Citral as monoterpene oxygenated derivative because they are the most relevant compounds of each class, while the sesquiterpenes are minority components and we did not predict their behaviour. The prediction of the Ai, Ci and k constants of Citral was carried out on the basis of their physico-chemical properties shown in Table 4.10. Table 4.10. Physico-chemical properties of Citral and Geranial. MW: molecular weight; Tb: boiling temperature; density; solubility parameter. MW (g/mol) Citral Geranial 152.24 152.24 Tb (K) (kg/m3) (MPa)1/2 502.20 502.20 893.00 893.00 19.58 19.47 Once the constants were obtained, the prediction of the VLE of the Lemon oil was achieved employing the equations [4.23] and [4.24] and considering the composition of the mixture. The experimental measurements were performed at 323.15 and 343.15 K in a pressure range between 81 and 130 bar. Table 4.11 shows the experimental and predicted values of solubility of the Lemon oil. The predicted data fitted accurately the experimental solubility, although as it was observed previously, deviation increased at high pressure. In this case, the average standard deviation was 4.6 kg Lemon oil/ m3 CO2. It is important to emphasise that the solubility of the Lemon oil could vary in comparison with its pure components (Limonene and Citral) because when a component takes part in a complex real mixture, the accompanying compounds affect the global solubility. This could be the reason of the underestimation of the predicted data. 87 Binary Systems Table 4.11. Comparison between the experimental and predicted values of solubility of Lemon oil in supercritical CO2 at 323.15 K and 343.15 K. P (MPa) 8.1 8.6 8.9 9.2 9.5 9.8 9.95 10.1 8.6 9.7 10.8 12 12.6 13 SLemonOil (kg/m3) Exp. SLemon Oil (kg/m3) Pred. Temperature: 323.15 K 3.039 3.944 4.048 5.513 4.335 6.624 5.201 8.322 7.069 10.537 11.435 13.089 33.591 14.497 51.417 15.621 Temperature: 343.15 K 3.151 3.832 3.850 6.461 6.833 10.859 14.133 17.695 22.789 21.904 36.494 25.021 Std. Dev. 0.640 1.036 1.618 2.206 2.453 1.170 13.502 25.312 0.481 1.846 2.847 2.518 0.626 8.113 Next, the prediction of the behaviour of Orange oil in CO 2 was studied since this is a complex multicomponent mixture of terpenoids. Monoterpenes content of orange peel oil depends on the origin of the oranges and in this work three different compositions were studied (named as A, B and C). In general, they presented from 98.26 to 98.77 wt % of monoterpenes, where the most common were Limonene (95.35 wt %) and Linalool (0.39 wt %) obtained from Budich and Brunner [85]. Decanal and Geranial were also presented in the composition of the orange oil, but in this work only Geranial was considered in the mixture since Decanal is not a terpenoid. As far as we know, VLE data of Geranial have not been studied yet although it is present in several plants and fruits [86-91]. According to the described procedure, the Ai, Ci and k constants for Geranial were predicted based on its physico-chemical properties shown in Table 4.9 and considering the composition of the mixture [85], the prediction of the VLE of three different types of orange oil was achieved. Figure 4.17. shows the composition of the mixtures CO 2 + orange peel oil (three types: A, B, C) determined experimentally (markers) and by means of the proposed prediction (lines) at 313 and 323 K and in a pressure range from 70 to 130 bar. In this case, the solubility of the ternary mixture (Limonene, Linalool, Geranial) was comparatively lower than that of the single components. 88 Chapter 4 30 120 25 Pressure (bar) 3 Solubility (kg /m ) 140 100 80 60 Type A T: 313.15 K T: 323.15 K 40 20 0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 140 15 10 0 20 40 60 80 100 120 Pressure (bar) 15 Type B 3 Solubility (kg/m ) 100 80 60 Type B T: 313.15 K T: 323.15 K 40 20 12 T: 313.15 K T: 323.15 K 9 6 3 0 0 xCO2 (wt. %) 20 Type C 25 3 100 80 Type C T: 313.15 K T: 323.15 K 40 20 0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 xCO2 (wt. %) 60 80 100 120 100 120 30 120 60 40 Pressure (bar) Solubility (kg/m ) Pressure (bar) 5 0 xCO2 (wt. %) 0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 Pressure (bar) T: 313.15 K T: 323.15 K 20 120 140 Type A 20 T: 313.15 K T: 323.15 K 15 10 5 0 0 20 40 60 80 Pressure (bar) Figure 4.17. Comparison between experimental solubility data [85] for three different types of orange oils (A,B and C) at 313.15 K (■) and 323.15 K (▲), and the prediction at 313.15 K (—) and 323.15 K (—). According to the prediction of VLE for multicomponent mixtures shown in Figure 4.17, which contain compounds described or not in the literature, the semi-empirical model proposed is validated and the procedure can be used to estimate the solubility of natural oils in CO2 in a range of temperature between 313 and 333 K and at from atmospheric pressure up to 12 MPa. 4.3.4. Correlation and prediction of terpenes/CO2 using Equations of State Although the explained methodology is very easy, simple and versatile and allows the prediction of VLE data from the knowledge of the composition of natural oils, cubic Equations of State (EoS) have been generally used in the prediction of terpenes behaviour. Particularly, Peng-Robinson equation of state (PR-EoS) has been widely applied for the prediction of the phase equilibrium of nonpolar component mixtures at high pressures [24, 27, 82, 85, 92]. The general expression to describe PR-EoS is shown below: [4.26] 89 Binary Systems where P is the pressure, R is the ideal gas constant, T is the temperature, V is the molar volume, a is a measure of the attractive forces between the molecules and b is related to the size of the molecules. These two parameters are defined as: [4.27] [4.28] [4.29] [4.30] [4.31] where TR is the reduced temperature and is the acentric factor. The application of PR-EoS to mixtures (CO2/terpene oils, in this case) requires the knowledge of the critical constants (Table 4.11) and the use of mixing rules which relate the properties of the pure components to the properties of the mixtures [93]. The Van der Waals mixing rules are the most widely used: [4.32] [4.33] In equations [4.32] and [4.33], xi or xj represents the mole fraction of the component i or j in the mixture, respectively; aij and bij are parameteres corresponding to pure component when i=j, but when i≠j, are called the unlike-interaction parameter, and can be calculated following the next expressions: [4.34] [4.35] where kij is known as the coupling or binary parameter and is obtained from the correlation of experimental data CO2/terpene. Generally, the values of k12 for the most known binary systems CO2/terpenoids are collected in the literature, but the interactions between the different compounds involved in a complex mixture at high pressure are difficult to achieve. In this section, the correlation of experimental VLE data using PR-EoS was studied in order to obtain the binary parameters of the mixture. The k12 between the mentioned terpenoids and CO2 was obtained from fitting the referenced data via Newton-Raphson nonlinear regression procedure using as objective function the maximum likelihood with a convergence tolerance of 0.0001. Their values together with their standard deviation have been shown in the Table 4.12. 90 Chapter 4 Table 4.12. Critical constants, acentric factor and binary interaction parameter obtained from the vapour-liquid equilibrium correlation of terpenoids in CO 2 in the specific range of pressure and temperature. Terpenoids Tc (K) Pc (bar) Carvone 724.90 36.14 0.502 p- Cymene 652.00 28.00 0.376 Eucalyptol 643.16 27.00 0.419 Linalool 630.50 24.20 0.748 Limonene 653.00 28.20 0.381 -Pinene 644.00 27.60 0.221 -Terpinene 661.00 28.00 0.376 Terpenoids Temperature (K) Pressure (bar) k12 Std. Dev. Ref. Carvone 303.55-324.05 18-96 0.099 0.028 [27] p- Cymene 303.15-343.15 1.9-110.7 0.074 0.054 This work Eucalyptol 318.2-323.2 77.5-98.0 0.083 [28] Linalool 313.18-333 7.8-110.6 0.058 [29] Limonene 313.78-333.25 8.3-110.0 0.083 0.003 This work -Pinene 313-333 62.9-109.3 0.072 0.011 [24] -Terpinene 313.15-343.15 4.9-110.5 0.085 This work It should be outlined that VLE data of some terpenoids has been widely studied and there are many different references in the literature. Generally, when there is an important amount of equilibrium data, dispersion of the binary interaction parameters could be observed. Nevertheless, on a regular basis, k 12 ranges between 0.058 and 0.099 when the working conditions of pressure are between 2 and 120 bar and temperature from 30 to 70 ºC, which fits pretty well with the typical values showed by the hydrocarbons/CO2 binary systems [94]. Figures 4.18-4.20 show the experimental (markers) and the PR-EoS correlated (line) data obtained in this work for p-Cymene, Limonene and -Terpinene. 91 Binary Systems 140 120 Pressure (bar) 100 T:30 ºC T:30 ºC Correlation T:50 ºC T:50 ºC Correlation T:70 ºC T:70 ºC Correlation 80 60 40 20 0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.990 0.995 1.000 x/y CO2 Figure 4.18. Equilibrium phase composition experimental data (markers) from this work and predicted for the CO2/p-Cymene binary system at 30, 50 and 70ºC. k 12:0.074. 140 120 T:30 ºC T:30 ºC Correlation T:40 ºC T:40 ºC Correlation Pressure (bar) 100 80 60 40 20 0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.990 0.995 1.000 x/y CO2 Figure 4.19. Equilibrium phase composition experimental data (markers) from this work and predicted for the CO2/Limonene binary system at 30 and 40ºC. k 12:0.083. 92 Chapter 4 140 120 Pressure (bar) 100 T:40ºC T:40ºC Correlation T:50ºC T:50ºC Correlation T:70ºC T:70ºC Correlation 80 60 40 20 0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.990 0.995 1.000 x/y CO2 Figure 4.20. Equilibrium phase composition experimental data (markers) from this work and predicted for the CO2/-Terpinene binary system at 40, 50 and 70ºC. k 12:0.085. In general, good agreement between the experimental and correlated vapour-liquid equilibrium data was achieved. The behaviour of the experimental data is well represented using the PR-EoS model with the optimal values of the binary interaction parameter shown. It is observed that at constant temperature, raising the pressure higher amounts of Terpene are transferred from the liquid to the vapour phase because the density of the supercritical fluid increases as well as its solvating power becomes greater. Thus, it can be concluded that the knowledge of the phase behaviour by semiempirical or theoretical methods would allow the determination of the most suitable working conditions to solubilise the terpenes in CO 2. But on a regular basis, at mild temperatures, pressures above 100 bar are enough to ensure that the selected terpene oil is in the vapour phase. 93 Binary Systems 4.4. Polystyrene/CO2 Results 4.4.1. Sorption of CO2 in Polystyrene. Plasticization The characteristic parameters for carbon dioxide and Polystyrene are tabulated in the literature [95, 96], but in the case of CO2, the mentioned parameters vary widely depending on the reference. Possibly it is due to the fact that different thermodynamic properties were used to obtain the parameters and different regions of pressure and temperature were studied in each case. By this reason, an initial screening of the most suitable characteristic parameters for CO2 was carried out to establish a reliable prediction of sorption. Table 4.13 gathers a set of different characteristic parameters shown in the literature. Table 4.13. Sanchez-Lacombe EOS characteristic parameters for carbon dioxide P* (bar) T*(K) *(kg/m3) Ref. 7203 f(t) 1580 [97] 4181 316 1369 [98] 6596 283 1620 [99] 5745 305 1510 [100] 7195 280 1618 [101] 5745 305 1510 [102] 5737 309 1504 [103] 4643 328 1426 [104] 4126 316 1369 [98] 6510 283 1620 [99] 5670 305 1510 [100] 7101 280 1618 [101] 6590 283 1620 [105] According to the characteristic parameters compile in Table 4.13, density of CO2 was calculated following SL-EoS and the values were compared with those obtained from the equation of Bender [106] and Span & Wagner [107] (Figure 4.21). 94 Chapter 4 1000 900 800 * Density (kg/m3) * * 3 P (bar) T (K) (kg/m ) Ref. 4181 316 1369 [99] 6596 283 1620 [100] 5745 305 1510 [101] 7195 280 1618 [102] 5745 305 1510 [103] 5737 309 1504 [104] 4643 328 1426 [105] 4126 316 1369 [99] 6510 283 1620 [100] 5670 305 1510 [101] 7101 280 1618 [102] 6590 283 1620 [106] Eq. of Span and Wagner [108] Eq. of Bender [107] 700 600 500 400 300 200 100 0 0 25 50 75 100 125 150 175 200 225 Pressure (bar) Figure 4.21. Influence of the characteristic parameters of CO 2 on its density. Regarding Figure 4.21, the values reported by Panayiotou and Sanchez [103] (P*=5737 bar; T*= 309 K; *= 1510 kg/m3) achieved the best fit among the studied parameters, although the supercritical phase transition is not pretty well fitted in any case. The mentioned characteristics parameters together with P*=3870 bar; T*= 7399 K; *= 1108 kg/m3 from Sato et al. [97] in the case of PS, were selected to apply SL-EoS. CO2 sorption in PS data vary mainly due to the different methods and experimental conditions used. Aubert [108] and Zhang et al. [51] measured the solubility using gravimetric methods by means of a quartz crystal microbalance while Sato et al. determined the sorption by the pressure-decay method [96] or by the magnetic suspension balance [97]. Also, some authors employed combination of two methods to be more accurate due to changes in the volume of the polymer. Thus, Hilic et al. [47] combined the magnetic suspension balance with the pressure-decay method. On the other hand, high temperature were generally used to melt the polymer (Tg ≈ 373 K), but the plasticizing effect of CO2 decreases dramatically the Tg and low operating temperatures can be used to perform the sorption experiments. The experimental data of CO2 sorption in Polystyrene at 30 ºC are depicted in Figure 4.22. The sorption of CO2 increases almost linearly with pressure up to 100 bar where it reaches a constant value close to 0.16 g CO2/ g PS. 95 Binary Systems 1000 0.15 800 0.10 600 400 0.05 0 Experimental data at T: 30 ºC Shieh and Liu [109] at T: 32 ºC Zhang et al. [89] at T: 35 ºC Aubert [108] at T: 40 ºC Pantoula and Panayiotou [72] at T: 51 ºC Hilic et al. [85] at T: 65 ºC 0.20 0.15 3 200 0.00 Sorption (gCO2/gPS) 1200 This work at T: 30 ºC Density of CO2 at T: 30 ºC Density of CO 2 (kg/m ) Sorption (gCO2/gPS) 0.20 0.10 0.05 0.00 0 50 100 150 200 250 300 350 Pressure (bar) Figure 4.22. (Upper) CO2 sorption in PS as a function of pressure at 30 ºC. The relationship between the sorption of CO2 and density can be observed in the right axis. (Below) Comparison between the experimental data (■) obtained in this work and those from the literature as it is indicated by the markers: (○) Shieh and Liu [109]; () Zhang et al. [51]; () Aubert [108]; (◊) Pantoula and Panayitou [34] and ( ) Hilic et al. [47]. A comparison between our experimental data and those reported in the literature at temperature between 32 and 65 ºC are shown in Figure 4.21. Data could be divided in two groups attending to the maximum sorption values. Shieh and Liu [109], Zhang et al. [51] and Pantoula and Panayiotou [34] data reached values of sorption around 0.06 g CO2/g PS at 32, 35 and 51 ºC, respectively. On the other hand, Aubert [108] and Hilic et al. [47] showed an increase on sorption reaching 0.16 g CO 2/g PS at 40 and 65 ºC, respectively. The last cited references present experimental data similar to those shown in Figure 4.22. According to the literature, when temperature increases, solubility of CO 2 decreases due to the density decrease [110]. This trend has been generally observed in many gas/polymer systems, where the highest values of sorption could be reached at high pressure and low temperature. By this reason, the sorption of CO2 is commonly 96 Chapter 4 related with its density, thus, when density increases higher amount of CO 2 are absorbed into the polymer. The sorption of CO2 in Polystyrene was correlated using Henry, Dual-Mode model equations and Sánchez-Lacombe EoS (Figure 4.23). The easiest model to correlate the sorption of CO2 into the PS follows the law of Henry (equation 4.1), where the sorption of CO2 is proportional to the pressure. It is especially suitable to fit the sorption of CO2 at low pressures. 0.20 0.18 Sorption (gCO2/gPS) 0.16 0.14 0.12 0.10 0.08 0.06 Experimental data Henry equation Dual-mode model Sánchez-Lacombe prediction 0.04 0.02 0.00 0 50 100 150 200 250 Pressure (bar) Figure 4.23. Correlation of sorption experimental data (■) using different models: Henry’s law (), Dual-mode model () and Sánchez-Lacombe Equation of State (). The good agreement between the experimental data and Sánchez-Lacombe EoS can be checked. Figure 4.23 compares the model prediction and correlation and the experimental data for the systems. It is observed that SL-EoS correlates the experimental data with reasonable accuracy and improves the results shown by the Dual-Mode model. At low pressure ( 75 bar) Henry’s law correlates pretty well the experimental results, while the dual-mode model and the SL-EoS overestimates the values of sorption. Nevertheless, when pressure increases dual-mode model and the SL-EoS enhance the correlation of the sorption of CO2 in PS. In contrast to Henry’s law that predicts a linear relation between the solubility and the saturation pressure, SLEoS can correlate the thermodynamic data with enhanced quality in a wider pressure range. It is observed that at pressure between 75 and 100 bar, the experimental values are higher than the correlated. Glassy polymers have nonequilibrium excess free volume and the CO2 could be absorbed into these microcavities achieving an increase in sorption [46]. 97 Binary Systems 4.4.2. Glass transition temperature The CO2 sorption into the PS is the responsible of the subsequent plasticization of the polymer. These molecular level phenomena are observed macroscopically by the decrease of the glass transition temperature (Tg) of the polymer [37]. Polymers in glassy state present a fixed free volume and any polymeric chain, but when temperature increases or a plasticizer is added, the free volume is expanded and the energy of molecular thermodynamic movement is enlarged promoting the movement of the polymeric chains [111]. Sorption of CO2 in PS lowers its Tg significantly below that observed at atmospheric pressure (Tg ≈ 373 K). In this case, the glass transition temperature of PS decrease is related with the increase of CO2 sorption. The influence of pressure on the CO 2 sorption and on the glass transition temperature of PS is shown in Figure 4.24. 120 0.18 110 0.16 100 0.14 80 0.12 Tg (ºC) 70 0.10 60 0.08 50 40 0.06 30 0.04 20 Glass transition Temperature (ºC) Sorption of CO2 in PS (g CO2 / g PS) 10 0 0 25 50 75 100 125 150 175 200 Sorption (gCO2/gPS) 90 0.02 0.00 225 Pressure (bar) Figure 4.24. Influence of pressure on the glass transition temperature of PS (●-left axis) and the sorption of CO2 (-■- right axis). The decrease of Tg is related with the increase of CO 2 sorption. Figure 4.24 shows an initial almost linear decrease of Tg at low pressures following the opposite trend to the sorption of CO2. Thus, the Tg of PS decreases with increasing the mass of CO2. When the critical pressure of CO 2 is reached, Tg remains approximately constant around 30ºC due to the saturation of PS. The plasticization of PS with compressed CO2 depends on many factors, such as the free volume of the polymer under pressure, the polymeric chains flexibility, the critical temperature and in particular, the sorption of CO2 in PS [41]. 98 Chapter 4 It is confirmed that the solubilisation of CO 2 into the polymer induces the plasticization of the plastic which is noticeable in a reduction of Tg [112, 113]. This effect could be predicted according to Chow’s equation [114]: [4.36] [4.37] [4.38] where Tg is the glass transition temperature of the polymer under pressure; Tg,0 is the glass transition temperature of the pure polymer at atmospheric pressure; is a nondimensional parameter defined in expression [4.39]; Mm is the molar mass of the monomeric unit which makes up the polymer (104.15 g/mol for PS) ; Md is the molar mass of the dissolved gas (44.01 g/mol for CO2); R is the gas constant; Cpp is the excess transition isobaric specific heat of the pure polymer (0.2593 J/g·K) and z is the lattice coordination number which can be either 1 or 2. Generally for polymer with small repeat units, as the PS, z is equal to 1 [100] but Chow proposed z=2 for the PS [114]. The use of Chow’s model becomes complicated due to the uncertainty of the estimation of z values. When z is two or greater, retrograde vitrification of PS is predicted, nevertheless, in the range of working conditions it was not observed experimentally. Furthermore, Wingert et al. [115] concluded that optimization of z is also dependant of the excess transition isobaric specific heat of the polymer. Due to the inexistence of retrograde vitrification in our data, in this work z=1 was selected according to Cao et al. [100]. To predict the Chow’s curve of Tg vs pressure, the calculation of concentration is required. Sánchez-Lacombe EoS (Figure 4.22) was used and the results of the model were compared with the values of Tg obtained experimentally (Figure 4.25). 110 100 90 80 Tg (ºC) 70 60 50 40 30 20 10 0 0 20 40 60 80 100 120 140 160 180 200 Pressure (bar) Figure 4.25. Prediction of the decrease of Tg of PS (●) using the Chow model. The perfect agreement between the experimental data and the prediction is observed. 99 Binary Systems The predicted data (solid curve) shown in Figure 4.25 fitted accurately the experimental values of the glass transition temperature of PS under CO 2 pressure, and small deviation was observed in the range of pressure studied. 4.4.3. Interfacial Tension The influence of CO2 on the Interfacial Tension (IFT) of PS was studied theoretically and results were compared with the literature [54, 55, 116]. It should be outlined that the molecular weight and crystallinity of PS plays a crucial role on the IFT and by this reason, important dispersion of data is shown in the literature. On a regular basis, IFT of PS decreases at rising pressure in CO 2 atmosphere and the slope of the IFT-curve decreases at elevated pressure. The decrease of IFT is attributed to two phenomena: when pressure increases, the free energy density of CO 2 becomes closer to that of the polymer phase and the second effect is related with the amount of CO 2 absorbed in the PS, which plasticizes the polymer [117]. The relationship between the sorption of CO2 and the IFT was described by Pastore Carbone et al. [117] using the empirical expression shown below: [4.39] where sol is the interfacial tension of the mixture PS/CO2, wCO2 is the weight fraction of the dissolved CO2, pol is the interfacial tension of the polymer and r is a fitting parameter. The estimation of wCO2 in the range where the IFT was measured experimentally was predicted according to S-L EoS. Initially, the agreement between the predicted wCO2 and the values shown in the literature [96, 97] in a wider range of temperature are depicted in Figure 4.26. 0.10 Experimental data from Sato et al. [96,97] at 423.15 K Sánchez-Lacombe prediction at 423.15 K Sorption (gCO2/gPS) 0.08 0.06 0.04 0.02 0.00 0 25 50 75 100 125 150 175 200 Pressure (bar) Figure 4.26. Prediction and comparison of the sorption of CO 2 in PS at 423.15 K to validate the SL-EoS at higher temperature. Experimental data were obtained from Sato et al. [96, 97]. 100 Chapter 4 The average standard deviation between experimental and predicted data was 0.006 gCO2/ gPS which means that sorption of CO2 in PS could be successfully predicted using S-L EoS in all the range of temperature and pressure. Thus, the predicted values of sorption were used to correlate the IFT of the mixture according to eq. [4.40] The IFT of PS (pol) was calculated following the Macleod-Sugden equation [118] and the effect of temperature and pressure on its volume was estimated according to the Tait-relation [59]. The fitted parameter r ranged between 0.34 and 0.37, nevertheless, the correlated values and the experimental data reported in the literature did not fit well, although the predicted values of sorption were reliable. This fact highlights the importance of the experimental measurement of the IFT of polymers at high pressure. This property is affected by the molecular weight or the crystallinity of the polymer which implies huge estimation uncertainty. 40 Temperature: 90 ºC from Jaeger et al. [92] Temperature: 160 ºC from Jaeger et al. [92] Temperature: 200 ºC from Li et al. [93] Temperature: 210 ºC from Li et al. [93] 35 30 (mN/m) 25 20 15 10 5 0 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 wCO2 40 Temp: 90 ºC from Jaeger et al. [92] Temp: 160 ºC from Jaeger et al. [92] Temp: 200 ºC from Li et al. [93] Temp: 210 ºC from Li et al. [93] 35 30 (mN/m) 25 20 15 10 5 0 0 20 40 60 80 100 120 140 160 180 200 Pressure (bar) Figure 4.27. Influence of sorption (upper figure) and pressure (figure below) on the decrease of the interfacial tension of PS. Experimental data are from Jaeger et al.[54] and Li et al. [55]. The influence of molecular weight distribution and crystallinity makes difficult the accurate prediction of the decrease of IFT. 101 Binary Systems The relationship between the sorption of CO 2 in PS and the decrease of the IFT is observed in Figure 4.27 (upper). Although a general trend is observed, the references studied different types of PS, which complicates the design of a general fitting independently of the kind of polymer used. The best agreement between experimental and predicted data was achieved using the general Macleod-Sugden [119] correlation for mixtures. The exponential factor (n) was fitted using the cited references and the optimum value was between 3.4 and 3.7 (Figure 4.27 down). 4.4.4. Viscosity The determination of viscosity at high pressure is still nowadays a challenge since the diluents must be retained in solution under working conditions of pressure and temperature [120]. CO2 has lower density and higher compressibility than pure polymer, so its dissolution into the PS results in swelling of the polymer. Consequently, an increase in free volume is produced and the transport properties such as viscosity can be enhanced [121]. Figure 4.28 shows a typical curve obtained from the rheometer measurements at 120 ºC and atmospheric pressure. 10 7 10 6 10 5 10 4 10 3 10 2 10 1 10 0 Viscosity (Pa·s) 120 ºC P atm 10 -1 1E-3 0.01 0.1 1 10 100 1000 Speed [1/min] Figure 4.28. Determination of the linear viscoelastic region of PS at 120 ºC and atmospheric pressure. The polymer shows Newtonian behaviour between 0.001 and 0.01 min-1. The error bars represent the standard deviation of the measurements from the mean. There are three different regions: a linear viscoelastic region (shear rate 0.01 min1), a depression of viscosity (above the critical shear rate) and finally the stabilization of the viscosity. During the initial stage, the viscosity is independent of shear rate and PS behaves as a Newtonian fluid. In the second stage the viscosity decrease suggests that the polymeric chains are disentangled in the direction of the shear. Finally, when the viscosity is kept constant at high shear rate implies the 102 Chapter 4 alienation of the chains [122]. PS exhibited a viscosity of 313.38 kPa·s at 120 ºC and 0.1 MPa, similar to the values showed in the literature by Wingert et al. [120] at 140 ºC. The effect of pressure and temperature on the viscosity of PS at a constant shear rate (0.01 min-1) was studied (Figure 4.29). The presence of CO2 (right axis) is the responsible of the viscosity reduction in the PS due to its plasticizing effect. Viscosity decreases significantly when CO 2 is absorbed in the polymer and an increase in the sorption entails a higher decrease of viscosity. 10 6 0.18 0.16 Viscosity (Pa·s) 0.14 5 0.12 10 4 10 3 10 0.10 0.08 2 0 25 50 75 CO2 Sorption at T: 30 ºC 0.06 Viscosity at T: 100 ºC Viscosity at T: 80 ºC Viscosity at T: 60 ºC Viscosity at T: 40 ºC 0.04 100 125 0.02 Sorption (g CO2/g PS) 10 0.00 150 Pressure (bar) Figure 4.29. Effect of pressure and temperature on the decrease of viscosity of PS in CO 2. The relationship between the drop of viscosity (left axis) and the sorption of CO 2 (right axis) can be observed. In general, in the standard systems, when pressure increases, viscosity raises; but in this case, the viscosity is decreased by the presence of CO 2. This small molecule is absorbed among the chains of the PS increasing the free volume and their mobility which causes the reduction in the viscosity [123]. To obtain reliable results, experiments were carried out when the sample was saturated with CO2. The changes in the viscosity along time allow the determination of the diffusion of CO2 in the PS. The sorption of CO2 in the polymer causes a decrease in viscosity, till the PS is saturated (Figure 4.30). 103 1.2x10 6 1.1x10 6 1.0x10 6 9.0x10 5 8.0x10 5 7.0x10 5 6.0x10 5 5.0x10 5 4.0x10 5 3.0x10 5 2.0x10 5 1.0x10 5 P: 100 bar; T: 60 ºC Viscosity (Pa·s) (R) 1.0 0.8 0.6 0.4 (R) Viscosity (Pa·s) Binary Systems 0.2 0.0 0 500 1000 1500 2000 2500 3000 3500 4000 0.0 4500 Time (s) Figure 4.30. Evolution of viscosity of PS as a function of time till saturation of polymer is reached, at 60 ºC and 100 bar. The initial point (left axis) is the viscosity of the polymer without CO 2 at 100 bar and 60 ºC and the points shown in the right axis represent the relative viscosity every 5 min. Relative viscosity R (t) is defined as: [4.40] where (t) is the viscosity at time t, (0) is the viscosity at atmospheric pressure, it means, when the concentration of CO 2 is zero and () is the viscosity at saturation. From experimental data shown in Figure 4.30, the diffusion coefficient of CO2 in PS was determined assuming that the gas is diffused into the sample through the upper side of the polymer disk, the diffusion coefficient is constant at each value of pressure and temperature and Fickian diffusion is accomplished [124]. The effect of pressure and temperature on the diffusion coefficient of CO 2 in PS is shown in Figure 4.31. 104 1E-9 1E-10 1E-10 2 1E-9 1E-11 1E-12 20 40 60 80 100 60 90 120 150 1E-12 180 1E-13 210 2 1E-13 1E-11 Pressure: 100 bar Pressure: 80 bar Pressure: 83.53 bar from Sato et al. [97] Pressure: 63.19 bar from Sato et al. [97] Pressure: 44.26 bar from Sato et al. [97] Pressure: 24.33 bar from Sato et al. [97] Temperature: 60 ºC Temperature: 80 ºC Temperature: 100 ºC Temperature: 100 ºC from Sato et al. [97] Temperature: 150 ºC from Sato et al. [97] Temperature: 200 ºC from Sato et al. [97] Mutual Diffusion Coefficient (m /s) Mutual Diffusion Coefficient (m /s) Chapter 4 Pressure (bar) Temperature (ºC) Figure 4.31. Influence of pressure (left figure) and temperature (right figure) on the mutual diffusion coefficients of CO2 in PS in a range of pressure from 20 to 100 bar and temperature from 40 to 100 ºC. Our experimental data were compared with the literature [97] and a good agreement between them was observed. The diffusion coefficient of CO2 in rubbery PS (above its Tg) is around 10-10 m2/s, typical value exhibited by high molecular weight polymers [52, 125]. From the comparison of the diffusion coefficient obtained in this work and the data from Sato et al. [97], the feasibility of the viscosity measurements to determine the diffusivity is confirmed. According to Figure 4.31, when pressure and temperature increase, the mutual diffusion coefficient increases, meaning that CO2 is diffuses faster at those conditions. The effect of temperature on the diffusion coefficients has been generally described attending Arrhenius or Williams-Landel-Ferry model [126], but due to the lack of experimental data, this fact could not be checked. To sum up, the sorption of CO2 in Polystyrene is the responsible of the swelling and plasticization of the polymer which changes its glass transition temperature, its interfacial tension and its viscosity. The inclusion of a small molecule, as CO 2 is, among the polymeric chains causes an increase of the free volume and enlarges the movement of the polymer chains. On the other hand, the insolubility of the polymer is checked and it is used for confirming that precipitation of the PS from a terpene solution should be easily performed. 105 Binary Systems References [1] J.M. Arandes, J. Ereña, M.J. Azkoiti, M. Olazar, J. Bilbao, Thermal recycling of polystyrene and polystyrene-butadiene dissolved in a light cycle oil, Journal of Analytical and Applied Pyrolysis, 70 (2003) 747-760. [2] Y. Zhang, S.K. Mallapragada, B. Narasimhan, Dissolution of waste plastics in biodiesel, Polymer Engineering and Science, 50 (2010) 863-870. [3] Y. Kodera, Y. Ishihara, T. Kuroki, S. Ozaki, Feedstock Recycling of Plastics, in: M. Müller-Hagedorn, H. Bockhorn (Eds.) Solvo-Cycle Process: AIST's New recycling process for used plastic foam using plastics-derived solvent, Karlshrue, 2005, pp. 217-222. [4] S.S. Kim, J. Kim, J.K. Jeon, Y.K. Park, C.J. Park, Non-isothermal pyrolysis of the mixtures of waste automobile lubricating oil and polystyrene in a stirred batch reactor, Renewable Energy, 54 (2013) 241-247. [5] F. Kerton, R. Marriott, Alternative Solvents for Green Chemistry, 2nd ed., 2013. [6] S. Shikata, T. Watanabe, K. Hattori, M. Aoyama, T. Miyakoshi, Dissolution of polystyrene into cyclic monoterpenes present in tree essential oils, Journal of Material Cycles and Waste Management, 13 (2011) 127-130. [7] K. Hattori, S. Shikata, R. Maekawa, M. Aoyama, Dissolution of polystyrene into p-cymene and related substances in tree leaf oils, Journal of Wood Science, 56 (2010) 169-171. [8] C. Gutiérrez, M.T. García, I. Gracia, A. De Lucas, J.F. Rodríguez, The selective dissolution technique as initial step for polystyrene recycling, Waste and Biomass Valorization, 4 (2013) 29-36. [9] K. Hattori, S. Naito, K. Yamauchi, H. Nakatani, T. Yoshida, S. Saito, M. Aoyama, T. Miyakoshi, Solubilization of polystyrene into monoterpenes, Advances in Polymer Technology, 27 (2008) 35-39. [10] M.T. García, G. Duque, I. Gracia, A. De Lucas, J.F. Rodríguez, Recycling extruded polystyrene by dissolution with suitable solvents, Journal of Material Cycles and Waste Management, 11 (2009) 2-5. [11] T. Noguchi, M. Miyashita, Y. Lnagaki, H. Watanabe, A new recycling system for expanded polystyrene using a natural solvent. Part 1. A new recycling technique, Packaging Technology and Science, 11 (1998) 19-27. [12] E. Breitmaier, Terpenes: Flavors, Fragrances, Pharmaca, Pheromones, Wiley, 2006. [13] C. Gutiérrez, M.T. García, I. Gracia, A. De Lucas, J.F. Rodríguez, A practical approximation to design a process for polymers recycling by dissolution, Una aproximación práctica para el diseño de un proceso de reciclado de polímeros mediante disolución, 68 (2011) 181-188. [14] F.W. Billmeyer, Textbook of polymer science, Wiley, New York, 1984. [15] C.M. Hansen, Hansen solubility parameters: a user's handbook, CRC Press, 2000. [16] C.D. Vaughan, Using solubility parameters in cosmetics formulation, Journal of the Society of Cosmetic Chemists of Japan, 36 (1985) 319-333. [17] S. Raeissi, S. Diaz, S. Espinosa, C.J. Peters, E.A. Brignole, Ethane as an alternative solvent for supercritical extraction of orange peel oils, Journal of Supercritical Fluids, 45 (2008) 306-313. [18] M. Sato, M. Goto, T. Hirose, Fractional extraction with supercritical carbon dioxide for the removal of terpenes from citrus oil, Industrial and Engineering Chemistry Research, 34 (1995) 3941-3946. 106 Chapter 4 [19] A.I. Cooper, Polymer synthesis and processing using supercritical carbon dioxide, Journal of Materials Chemistry, 10 (2000) 207-234. [20] P.T. Anastas, W. Leitner, P.G. Jessop, Handbook of Green Chemistry, Green Solvents, Supercritical Solvents, Wiley, 2013. [21] S.G. Kazarian, Polymer processing with supercritical fluids, Polymer Science Series C, 42 (2000) 78-101. [22] M. Drescher, O. Seidel, D. Geanǎ, High pressure vapor-liquid equilibria in the ternary system orange peel oil (limonene) + ethanol + carbon dioxide, Journal of Supercritical Fluids, 23 (2002) 103-111. [23] I. Gracia, M.T. García, J.F. Rodríguez, M.P. Fernández, A. de Lucas, Modelling of the phase behaviour for vegetable oils at supercritical conditions, Journal of Supercritical Fluids, 48 (2009) 189-194. [24] M. Akgün, N.A. Akgün, S. Dinçer, Phase behaviour of essential oil components in supercritical carbon dioxide, Journal of Supercritical Fluids, 15 (1999) 117-125. [25] Y.J. Sheng, P.C. Chen, Y.P. Chen, D.S.H. Wong, Calculations of solubilities of aromatic compounds in supercritical carbon dioxide, Industrial and Engineering Chemistry Research, 31 (1992) 967-973. [26] F.E. Wubbolts, O.S.L. Bruinsma, G.M. Van Rosmalen, Measurement and modelling of the solubility of solids in mixtures of common solvents and compressed gases, Journal of Supercritical Fluids, 32 (2004) 79-87. [27] T. Gamse, R. Marr, High-pressure phase equilibria of the binary systems carvone-carbon dioxide and limonene-carbon dioxide at 30, 40 and 50ºC, Fluid Phase Equilibria, 171 (2000) 165-174. [28] H.A. Matos, E.G. De Azevedo, P.C. Simoes, M.T. Carrondo, M.N. Da Ponte, Phase equilibria of natural flavours and supercritical solvents, Fluid Phase Equilibria, 52 (1989) 357-364. [29] C.-M.J. Chang, C.-C. Chen, High-pressure densities and P-T-x-y diagrams for carbon dioxide+linalool and carbon dioxide+limonene, Fluid Phase Equilibria, 163 (1999) 119-126. [30] J.d.C. Francisco, B. Sivik, Solubility of three monoterpenes, their mixtures and eucalyptus leaf oils in dense carbon dioxide, The Journal of Supercritical Fluids, 23 (2002) 11-19. [31] S.D. Yeo, E. Kiran, High-pressure density and viscosity of polystyrene solutions in methylcyclohexane, Journal of Supercritical Fluids, 15 (1999) 261-272. [32] H.M. Woods, M.M.C.G. Silva, C. Nouvel, K.M. Shakesheff, S.M. Howdle, Materials processing in supercritical carbon dioxide: Surfactants, polymers and biomaterials, Journal of Materials Chemistry, 14 (2004) 1663-1678. [33] O.R. Davies, A.L. Lewis, M.J. Whitaker, H. Tai, K.M. Shakesheff, S.M. Howdle, Applications of supercritical CO2 in the fabrication of polymer systems for drug delivery and tissue engineering, Advanced Drug Delivery Reviews, 60 (2008) 373387. [34] M. Pantoula, C. Panayiotou, Sorption and swelling in glassy polymer/carbon dioxide systems: Part I. Sorption, Journal of Supercritical Fluids, 37 (2006) 254-262. [35] I. Kikic, Polymer-supercritical fluid interactions, Journal of Supercritical Fluids, 47 (2009) 458-465. [36] O. Hölck, M. Böhning, M. Heuchel, M.R. Siegert, D. Hofmann, Gas sorption isotherms in swelling glassy polymers-Detailed atomistic simulations, Journal of Membrane Science, 428 (2013) 523-532. 107 Binary Systems [37] I. Kikic, F. Vecchione, P. Alessi, A. Cortesi, F. Eva, N. Elvassore, Polymer plasticization using supercritical carbon dioxide: Experiment and modeling, Industrial and Engineering Chemistry Research, 42 (2003) 3022-3029. [38] P. Alessi, A. Cortesi, I. Kikic, F. Vecchione, Plasticization of polymers with supercritical carbon dioxide: Experimental determination of glass-transition temperatures, Journal of Applied Polymer Science, 88 (2003) 2189-2193. [39] A. Bos, I.G.M. Pünt, M. Wessling, H. Strathmann, CO2-induced plasticization phenomena in glassy polymers, Journal of Membrane Science, 155 (1999) 67-78. [40] J.R. Royer, Y.J. Gay, J.M. Desimone, S.A. Khan, High-pressure rheology of polystyrene melts plasticized with CO 2: Experimental measurement and predictive scaling relationships, Journal of Polymer Science, Part B: Polymer Physics, 38 (2000) 3168-3180. [41] L. Yu, H. Liu, L. Chen, Thermal behaviors of polystyrene plasticized with compressed carbon dioxide in a sealed system, Polymer Engineering & Science, 49 (2009) 1800-1805. [42] C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. De Lucas, M.T. García, Development of a strategy for the foaming of polystyrene dissolutions in scCO2, Journal of Supercritical Fluids, 76 (2013) 126-134. [43] C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. de Lucas, M.T. García, Preparation and characterization of polystyrene foams from limonene solutions, The Journal of Supercritical Fluids, 88 (2014) 92-104. [44] H.E. Park, J.M. Dealy, Effects of supercritical fluids, pressure, temperature, and molecular structure on the rheological properties of molten polymers, in, Monterey, CA, 2008, pp. 366-368. [45] T. Spyriouni, G.C. Boulougouris, D.N. Theodorou, Prediction of sorption of CO2 in glassy atactic polystyrene at elevated pressures through a new computational scheme, Macromolecules, 42 (2009) 1759-1769. [46] S. Kanehashi, K. Nagai, Analysis of dual-mode model parameters for gas sorption in glassy polymers, Journal of Membrane Science, 253 (2005) 117-138. [47] S. Hilic, S.A.E. Boyer, A.A.H. Pádua, J.P.E. Grolier, Simultaneous measurement of the solubility of nitrogen and carbon dioxide in polystyrene and of the associated polymer swelling, Journal of Polymer Science, Part B: Polymer Physics, 39 (2001) 2063-2070. [48] I.C. Sanchez, R.H. Lacombe, Statistical thermodynamics of polymer solutions, Macromolecules, 11 (1978) 1145-1156. [49] M. Pantoula, J. von Schnitzler, R. Eggers, C. Panayiotou, Sorption and swelling in glassy polymer/carbon dioxide systems. Part II-Swelling, Journal of Supercritical Fluids, 39 (2007) 426-434. [50] L.N. Nikitin, M.O. Gallyamov, R.A. Vinokur, A.Y. Nikolaec, E.E. Said-Galiyev, A.R. Khokhlov, H.T. Jespersen, K. Schaumburg, Swelling and impregnation of polystyrene using supercritical carbon dioxide, Journal of Supercritical Fluids, 26 (2003) 263-273. [51] Y. Zhang, K.K. Gangwani, R.M. Lemert, Sorption and swelling of block copolymers in the presence of supercritical fluid carbon dioxide, Journal of Supercritical Fluids, 11 (1997) 115-134. [52] C. Tsioptsias, C. Panayiotou, Simultaneous determination of sorption, heat of sorption, diffusion coefficient and glass transition depression in polymer-CO2 systems, Thermochimica Acta, 521 (2011) 98-106. 108 Chapter 4 [53] A. Kasturirangan, C.A. Koh, A.S. Teja, Glass-transition temperatures in CO 2 + polymer systems: Modeling and experiment, Industrial and Engineering Chemistry Research, 50 (2011) 158-162. [54] P.T. Jaeger, R. Eggers, H. Baumgartl, Interfacial properties of high viscous liquids in a supercritical carbon dioxide atmosphere, Journal of Supercritical Fluids, 24 (2002) 203-217. [55] H. Li, L.J. Lee, D.L. Tomasko, Effect of Carbon Dioxide on the Interfacial Tension of Polymer Melts, Industrial and Engineering Chemistry Research, 43 (2004) 509-514. [56] S.P. Nalawade, V.H.J. Nieborg, F. Picchioni, L.P.B.M. Janssen, Prediction of the viscosity reduction due to dissolved CO 2 of and an elementary approach in the supercritical CO2 assisted continuous particle production of a polyester resin, Powder Technology, 170 (2006) 143-152. [57] P.A. Small, Some factors affecting the solubility of polymers, Journal of Applied Chemistry, 3 (1953) 71-80. [58] A.F.M. Barton, CRC handbook of solubility parameters and other cohesion parameters, CRC Press, 1991. [59] D.W. van Krevelen, K. te Nijenhuis, Properties of Polymers: Their Correlation with Chemical Structure; their Numerical Estimation and Prediction from Additive Group Contributions, Elsevier Science, 2009. [60] K.L. Hoy, Solubility parameter as a design parameter for water-borne polymers and coatings, Journal of Coated Fabrics, (1989) 53-67. [61] J. Szydlowski, L.P. Rebelo, H. Wilczura, W.A. Van Hook, Y. Melnichenko, G.D. Wignall, Liquid-liquid demixing from polystyrene solutions. Studies on temperature and pressure dependences using dynamic light scattering and neutron scattering, Fluid Phase Equilibria, 150 (1998) 687-694. [62] G. Bernardo, D. Vesely, Equilibrium solubility of alcohols in polystyrene attained by controlled diffusion, European Polymer Journal, 43 (2007) 938-948. [63] A. Güner, The algorithmic calculations of solubility parameter for the determination of interactions in dextran/certain polar solvent systems, European Polymer Journal, 40 (2004) 1587-1594. [64] M.T. García, I. Gracia, G. Duque, A.d. Lucas, J.F. Rodríguez, Study of the solubility and stability of polystyrene wastes in a dissolution recycling process, Waste Management, 29 (2009) 1814-1818. [65] P. Subra, P. Jestin, Screening design of experiment (DOE) applied to supercritical antisolvent process, Industrial and Engineering Chemistry Research, 39 (2000) 4178-4184. [66] I.H. Lin, P.F. Liang, C.S. Tan, Preparation of polystyrene/poly(methyl methacrylate) blends by compressed fluid antisolvent technique, Journal of Supercritical Fluids, 51 (2010) 384-398. [67] J. Holten-Andersen, K. Eng, Activity coefficients in polymer solutions, Progress in Organic Coatings, 16 (1988) 77-97. [68] T. Lindvig, M.L. Michelsen, G.M. Kontogeorgis, A Flory-Huggins model based on the Hansen solubility parameters, Fluid Phase Equilibria, 203 (2002) 247-260. [69] W.J. Cooper, P.D. Krasicky, F. Rodriguez, Effects of molecular weight and plasticization on dissolution rates of thin polymer films, Polymer, 26 (1985) 10691072. [70] K. Ueberreiter, The Solution Process, (1968) 219-257. [71] B.E. Poling, J.M. Prausnitz, J.P. O'Connell, Properties of Gases and Liquids (5th Edition), in, McGraw-Hill. 109 Binary Systems [72] J.S. Papanu, D.W. Hess, D.S. Soane, A.T. Bell, Dissolution of thin poly(methyl methacrylate) films in ketones, binary ketone/alcohol mixtures, and hydroxy ketones, Journal of the Electrochemical Society, 136 (1989) 3077-3083. [73] M. Okubo, H. Ahmad, Synthesis of temperature-sensitive submicron-size composite polymer particles, Colloid & Polymer Science, 273 (1995) 817-821. [74] P.J. Flory, Thermodynamics of high polymer solutions, The Journal of Chemical Physics, 9 (1941) 660-661. [75] M.L. Huggins, Solutions of long chain compounds, The Journal of Chemical Physics, 9 (1941) 440. [76] E. Eastwood, S. Viswanathan, C.P. O'Brien, D. Kumar, M.D. Dadmun, Methods to improve the properties of polymer mixtures: Optimizing intermolecular interactions and compatibilization, Polymer, 46 (2005) 3957-3970. [77] D.Z. Icoz, J.L. Kokini, Examination of the validity of the Flory-Huggins solution theory in terms of miscibility in dextran systems, Carbohydrate Polymers, 68 (2007) 59-67. [78] G. Ovejero, P. Pérez, M.D. Romero, I. Guzmán, E. Díez, Solubility and Flory Huggins parameters of SBES, poly(styrene-b-butene/ethylene-b-styrene) triblock copolymer, determined by intrinsic viscosity, European Polymer Journal, 43 (2007) 1444-1449. [79] A.R. Imre, W.A. Van Hook, The effect of branching of alkanes on the liquidliquid equilibrium of oligostyrene/alkane systems, Fluid Phase Equilibria, 187-188 (2001) 363-372. [80] J. Chrastil, Solubility of solids and liquids in supercritical gases, Journal of Physical Chemistry, 86 (1982) 3016-3021. [81] A. Cháfer, A. Berna, J.B. Montón, A. Mulet, High Pressure Solubility Data of the System Limonene + Linalool + CO 2, Journal of Chemical & Engineering Data, 46 (2001) 1145-1148. [82] S.A.B. Vieira de Melo, G.M.N. Costa, A.M.C. Uller, F.L.P. Pessoa, Modeling high-pressure vapor-liquid equilibrium of limonene, linalool and carbon dioxide systems, Journal of Supercritical Fluids, 16 (1999) 107-117. [83] M. Kondo, N. Akgun, M. Goto, A. Kodama, T. Hirose, Semi-batch operation and countercurrent extraction by supercritical CO2 for the fractionation of lemon oil, Journal of Supercritical Fluids, 23 (2002) 21-27. [84] F. Gironi, M. Maschietti, Phase equilibrium of the system supercritical carbon dioxide-lemon essential oil: New experimental data and thermodynamic modelling, Journal of Supercritical Fluids, 70 (2012) 8-16. [85] M. Budich, G. Brunner, Vapor–liquid equilibrium data and flooding point measurements of the mixture carbon dioxide+orange peel oil, Fluid Phase Equilibria, 158 (1999) 759-773. [86] M.C. Mesomo, M.L. Corazza, P.M. Ndiaye, O.R. Dalla Santa, L. Cardozo, A.D.P. Scheer, Supercritical CO2 extracts and essential oil of ginger (Zingiber officinale R.): Chemical composition and antibacterial activity, Journal of Supercritical Fluids, 80 (2013) 44-49. [87] B. Marongiu, A. Piras, S. Porcedda, E. Tuveri, Comparative analysis of the oil and supercritical CO2 extract of Cymbopogon citratus Stapf, Natural Product Research, 20 (2006) 455-459. [88] L.H.C. Carlson, R.A.F. Machado, C.B. Spricigo, L.K. Pereira, A. Bolzan, Extraction of lemongrass essential oil with dense carbon dioxide, Journal of Supercritical Fluids, 21 (2001) 33-39. 110 Chapter 4 [89] L.C. Argenta, J.P. Mattheis, X. Fan, F.L. Finger, Production of volatile compounds by Fuji apples following exposure to high CO2 or low O2, Journal of Agricultural and Food Chemistry, 52 (2004) 5957-5963. [90] C. Besada, G. Sanchez, A. Salvador, A. Granell, Volatile compounds associated to the loss of astringency in persimmon fruit revealed by untargeted GC-MS analysis, Metabolomics, 9 (2013) 157-172. [91] Š.I. Kirbalar, A. Gök, F.G. Kirbašlar, S. Tepe, Volatiles in Turkish clementine (Citrus clementina Hort.) peel, Journal of Essential Oil Research, 24 (2012) 153-157. [92] A.L. Oliveira, V.L.S. Melo, A.R. Guimaraes, F.A. Cabral, Modelling of highpressure phase equilibrium in systems of interest in the food engineering field using the peng-robinson equation of state with two different mixing rules, Journal of Food Process Engineering, 33 (2010) 101-116. [93] B.E. Poling, J.M. Prausnitz, J.P. O'Connell, The properties of gases and liquids, McGraw-Hill, New York, 2001. [94] E.A. Turek, R.S. Metcalfe, L. Yarborough, R.L. Robinson Jr, PHASE EQUILIBRIA IN CO2 -MULTICOMPONENT HYDROCARBON SYSTEMS: EXPERIMENTAL DATA AND AN IMPROVED PREDICTION TECHNIQUE, Society of Petroleum Engineers journal, 24 (1984) 308-324. [95] E. Kiran, Y. Xiong, W. Zhuang, Modeling polyethylene solutions in near and supercritical fluids using the Sanchez-Lacombe model, The Journal of Supercritical Fluids, 6 (1993) 193-203. [96] Y. Sato, M. Yurugi, K. Fujiwara, S. Takishima, H. Masuoka, Solubilities of carbon dioxide and nitrogen in polystyrene under high temperature and pressure, Fluid Phase Equilibria, 125 (1996) 129-138. [97] Y. Sato, T. Takikawa, S. Takishima, H. Masuoka, Solubilities and diffusion coefficients of carbon dioxide in poly(vinyl acetate) and polystyrene, The Journal of Supercritical Fluids, 19 (2001) 187-198. [98] R. Hariharan, B.D. Freeman, R.G. Carbonell, G.C. Sarti, Equation of state predictions of sorption isotherms in polymeric materials, Journal of Applied Polymer Science, 50 (1993) 1781-1795. [99] D.S. Pope, I.C. Sanchez, W.J. Koros, G.K. Fleming, Statistical thermodynamic interpretation of sorption/dilation behavior of gases in silicone rubber, Macromolecules, 24 (1991) 1779-1783. [100] G.P. Cao, T. Liu, G.W. Roberts, Predicting the effect of dissolved carbon dioxide on the glass transition temperature of poly(acrylic acid), Journal of Applied Polymer Science, 115 (2010) 2136-2143. [101] P.K. Kilpatrick, S.H. Chang, Saturated phase equilibria and parameter estimation of pure fluids with two lattice-gas models, Fluid Phase Equilibria, 30 (1986) 49-56. [102] M.N. Lotfollahi, H. Modarress, B. Khodakarami, VLE Predictions of strongly non-ideal binary mixtures by modifying Van der Waals and Orbey-Sandler mixing rules, Iranian Journal of Chemistry and Chemical Engineering, 26 (2007) 73-80. [103] C. Panayiotou, I.C. Sanchez, Swelling of network structures, Polymer, 33 (1992) 5090-5093. [104] A. Garg, E. Gulari, C.W. Manke, Thermodynamics of polymer melts swollen with supercritical gases, Macromolecules, 27 (1994) 5643-5653. [105] X. Wang, I.C. Sanchez, Welding immiscible polymers with a supercritical fluid, Langmuir, 23 (2007) 12192-12195. [106] E. Bender, Equations of state for ethylene and propylene, Cryogenics, 15 (1975) 667-673. 111 Binary Systems [107] R. Span, W. Wagner, A new equation of state for carbon dioxide covering the fluid region from the triple-point temperature to 1100 K at pressures up to 800 MPa, Journal of Physical and Chemical Reference Data, 25 (1996) 1509-1596. [108] J.H. Aubert, Solubility of carbon dioxide in polymers by the quartz crystal microbalance technique, Journal of Supercritical Fluids, 11 (1998) 163-172. [109] Y.T. Shieh, K.H. Liu, The effect of carbonyl group on sorption of CO 2 in glassy polymers, Journal of Supercritical Fluids, 25 (2003) 261-268. [110] Y. Sato, K. Fujiwara, T. Takikawa, Sumarno, S. Takishima, H. Masuoka, Solubilities and diffusion coefficients of carbon dioxide and nitrogen in polypropylene, high-density polyethylene, and polystyrene under high pressures and temperatures, Fluid Phase Equilibria, 162 (1999) 261-276. [111] X.K. Li, G.P. Cao, L.H. Chen, R.H. Zhang, H.L. Liu, Y.H. Shi, Study of the anomalous sorption behavior of CO2 into poly(methyl methacrylate) films in the vicinity of the critical pressure and temperature using a quartz crystal microbalance (QCM), Langmuir, 29 (2013) 14089-14100. [112] A.G. Wonders, D.R. Paul, Effect of CO2 exposure history on sorption and transport in polycarbonate, Journal of Membrane Science, 5 (1979) 63-75. [113] S.A. Stern, A.H. De Meringo, Solubility of Carbon Dioxide in cellulose acetate at elevated pressures, J Polym Sci Polym Phys Ed, 16 (1978) 735-751. [114] T.S. Chow, Molecular interpretation of the glass transition temperature of polymer-diluent systems, Macromolecules, 13 (1980) 362-364. [115] K. Liu, F. Schuch, E. Kiran, High-pressure viscosity and density of poly(methyl methacrylate) + acetone and poly(methyl methacrylate) + acetone + CO 2 systems, Journal of Supercritical Fluids, 39 (2006) 89-101. [116] H. Park, R.B. Thompson, N. Lanson, C. Tzoganakis, C.B. Park, P. Chen, Effect of temperature and pressure on surface tension of polystyrene in supercritical carbon dioxide, Journal of Physical Chemistry B, 111 (2007) 3859-3868. [117] M.G. Pastore Carbone, E. Di Maio, G. Scherillo, G. Mensitieri, S. Iannace, Solubility, mutual diffusivity, specific volume and interfacial tension of molten PCL/CO2 solutions by a fully experimental procedure: Effect of pressure and temperature, Journal of Supercritical Fluids, 67 (2012) 131-138. [118] R.C. Reid, J.M. Prausnitz, B.E. Poling, The properties of gases and liquids, 1987. [119] D.B. Macleod, On a relation between surface tension and density, Transactions of the Faraday Society, 19 (1923) 38-41. [120] M.J. Wingert, S. Shukla, K.W. Koelling, D.L. Tomasko, L.J. Lee, Shear viscosity of CO2-plasticized polystyrene under high static pressures, Industrial and Engineering Chemistry Research, 48 (2009) 5460-5471. [121] J.E. Mark, Physical properties of polymers handbook, AIP Press, 1996. [122] C.A. Kelly, S.M. Howdle, K.M. Shakesheff, M.J. Jenkins, G.A. Leeke, Viscosity studies of poly(DL -lactic acid) in supercritical CO2, Journal of Polymer Science, Part B: Polymer Physics, 50 (2012) 1383-1393. [123] C. Gutiérrez, M.T. Garcia, S. Curia, S.M. Howdle, J.F. Rodriguez, The effect of CO2 on the viscosity of polystyrene/limonene solutions, Journal of Supercritical Fluids, 88 (2014) 26-37. [124] M.H. Klopffer, B. Flaconnèche, Transport properties of gases in polymers: Bibliographic review, Transport de molécule gazeuses dans les polymères: Revue bibliograpique, 56 (2001) 223-244. 112 Chapter 4 [125] K. Taki, K. Nitta, S.I. Kihara, M. Ohshima, CO2 foaming of poly(ethylene glycol)/polystyrene blends: Relationship of the blend morphology, CO 2 mass transfer, and cellular structure, Journal of Applied Polymer Science, 97 (2005) 1899-1906. [126] G. Bernardo, R.P. Choudhury, H.W. Beckham, Diffusivity of small molecules in polymers: Carboxylic acids in polystyrene, Polymer, 53 (2012) 976-983. 113 Binary Systems 114 Chapter 5 TERNARY SYSTEMS: EQUILIBRIUM Chapter 5 deals about the study of the ternary mixtures CO 2/terpene oils/PS. Although the individual knowledge of the binary solutions provides a good idea regarding the global process, the determination of the phases’ equilibrium of the mixture is crucial for the accurate design. But also, the influence of CO 2 on the properties of the solutions is a key factor since the sorption of CO 2 plasticises the mixture changing dramatically their physic-chemical and rheological characteristics. According to this, the Chapter has been divided in three different sections: the study of CO2/Limonene/PS equilibrium, the determination of CO 2/pCymene/PS equilibrium and the measurement of viscosity and interfacial tension of both mixtures. Equilibrium of Ternary Mixtures 116 Chapter 5 Based on: C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. de Lucas, M.T. García, High-pressure phase equilibria of Polystyrene dissolutions in Limonene in presence of CO 2, The Journal of Supercritical Fluids, 84 (2013) 211-220. C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. de Lucas, M.T. García, Determination of the High-pressure phase equilibria of Polystyrene/p-Cymene in presence of CO2, The Journal of Supercritical Fluids, enviado. Graphical abstract Polystyrene Two-phases Terpene CO2 117 Equilibrium of Ternary Mixtures RESUMEN La disolución de residuos de Poliestireno en aceites terpénicos es una excelente alternativa para su reciclaje. En concreto, se seleccionaron el Limoneno y el pCimeno por tratarse de los disolventes en los que el polímero presentaba una elevada solubilidad, además de ser compuestos naturales, no tóxicos y estar disponibles comercialmente a un bajo precio. A continuación, mediante CO 2 a alta presión es posible retirar el disolvente para que precipite el polímero, ya que tanto el Limoneno como el p-Cimeno son muy solubles en CO2 y por el contrario, el Poliestireno presenta una total insolubilidad. Para definir las condiciones de operación más adecuadas a las que realizar la separación de los componentes, es fundamental el conocimiento del equilibrio de fases de la mezcla ternaria (CO2/terpeno/PS). En este Capítulo, se determina la solubilidad del Limoneno y del p-Cimeno en la fase rica en CO2. Para ello, se llevaron a cabo los experimentos en una celda provista de visores de volumen variable haciendo uso del método estático. Las condiciones a las que se realizaron las medidas experimentales fueron en un rango de temperatura de 25 a 40 ºC, concentración de la disolución inicial entre 0.05 y 0.80 g PS/ml terpeno y a presiones de hasta 150 bar. A partir de estos datos, se concluyó que las condiciones más adecuadas para conseguir la precipitación del Poliestireno a partir de su disolución en Limoneno o p-Cimeno son aquellas en las que se alcancen altos valores de densidad (alta presión y baja temperatura) y valores moderados de concentración. 118 Chapter 5 ABSTRACT Dissolution with terpenic solvents is presented as an alternative and original route to recycle Polystyrene wastes at room temperature. Limonene and p-Cymene were the selected solvent to perform the dissolution. The mentioned terpenes present high compatibility with Polystyrene besides being natural, non toxic and relatively low cost. The solvent removal is possible thanks to supercritical CO 2 since it provides high solubility of Limonene and p-Cymene and complete PS insolubility at moderated pressures and temperature. In order to determine the proper working conditions to conduct the precipitation of the polymer, accurate knowledge of the phase equilibrium for the ternary mixtures should be known. In this section, the solubility of Limonene and p-Cymene in the ternary system was determined. The phase equilibrium experiments were conducted in a variablevolume view cell employing the static method. These experiments were carried out in the temperature range of 25-40 ºC, at pressures up to 150 bar and in the concentration range of 0.05-0.80 g PS/ml Terpene. The optimum conditions to carry out the precipitation process of Polystyrene were predicted on the basis of the calculated values. The most suitable conditions to recycle Polystyrene were reached at high values of density, it means, high pressure, low temperature and moderated concentration. 119 Equilibrium of Ternary Mixtures 120 Chapter 5 5.1. Background of the Phases’ equilibrium The use of carbon dioxide as a non-solvent for polymer phase separation has been extensively reported in the literature [1-3]. It is based on the use of two liquid solvents that are completely miscible. The polymer is soluble in the first solvent, but not soluble in the second solvent. Therefore, the addition of the antisolvent induces the formation of a solution of the two liquids and the supersaturation and precipitation of the solute. This process is a powerful tool for solvent recovery in solution polymerisation processes as well as for separation or precipitation of polymers [4-6]. The ability of scCO2 to tune physico-chemical properties near its critical point (Tc: 31.1 ºC, Pc: 73.8 bar) provides a high diffusivity, high solvent power and the absence of surface tension which means higher phase separation velocities [7]. By the reported reasons, CO2 was selected as antisolvent in the recycling process of Polystyrene from its solution in terpenes. In this process, CO2 is miscible with the terpenes, but it acts as antisolvent for the polymer because its solubility in the CO 2 is negligible [3]. The polymer selected in this work presented high molecular weight and consequently it was insoluble in CO2 [8, 9], but terpene oils present low molecular weight and non polarity and by these reasons they are fully soluble in CO2 at mild operating conditions (Chapter 4). Due to the small size of the CO2 molecule, two-way diffusion is much faster than in the case of conventional liquid antisolvent. Thus, during the antisolvent precipitation process, the accelerated mass transfer between the mediums can facilitate very rapid phase separation and hence the production of materials with a well defined structure [1]. The knowledge of phase behaviour of the solute + solvent + antisolvent system plays an important role to determine suitable operating conditions to conduct precipitation of polymer and understanding the kinetic nucleation mechanism and growth of the particles. The control of products morphology is possible since CO2 modifies the properties of polymers as a consequence of its sorption and swelling and a good control of size and size distribution of the material is expected. The morphology of the spheres, foams or fibres is related to the location of phase separation [2]. Although many parameters should be considered during the optimization of these processes, the most important is the phase behaviour of the systems. A thermodynamic study of the system solute-solvent-antisolvent is extremely useful to address the feasibility of the process and to exploit the effects of pressure and temperature because it governs the mass transport [10]. Although the binary equilibriums have been described and they can be predicted based on the experimental data compilated in the literature (Chapter 4), they are not enough to determine the behaviour of the mixtures. The behaviour of the binary systems solvent-antisolvent could be dramatically modified because of the addition of a solute, especially when it concerns a polymer. Due to the presence of polymers in the process, equilibrium data and accurate predictions of the phases behaviour 121 Equilibrium of Ternary Mixtures are very difficult to achieve. Thus experimental work is needed because it will complete and improve the existing models. The potential applications of polymer in the chemical industries have generated a corpus of literature available on the scCO2 and polymer processing techniques. However, most of the studies using polymers and supercritical technology are carried out at high pressure and temperature under which the polymers are molten but consequently they can be degraded, changing the polystyrene morphology [1114]. As far as we know, few works provide phase equilibrium data of polymer solutions at low temperatures but it is important since polymer does not undergo chain degradation and a lower energetic consumption is required [15]. In this Chapter, we present the phase equilibrium data for the ternary systems CO2/Limonene/Polystyrene and CO2/p-Cymene/Polystyrene. The limits of the regions where an antisolvent process could be carried out were determined in order to select the most suitable working conditions. The influence of the pressure, the temperature and the initial concentration of the solution was evaluated and correlated. Also the ternary diagrams were built to highlight the importance of understanding high-pressure phase behaviour. Finally, in order to drawn some general trends, the solubility and sorption data were fitted to semi-empirical equations which allowed a global comprehension of the solubility of p-Cymene and CO2 in the ternary system. 5.2. The ternary system CO2/Limonene/PS In this section, phase equilibrium data for the ternary systems CO 2 (1) + Limonene (2) + PS (3) at different concentrations of PS in the organic solvent and at different temperatures and pressures are reported. With the aim of separate the PS from its solution in Limonene, it is necessary to determine when the homogeneous and/or heterogeneous regions are presented. Since working conditions of the separation operation are suited to ensure an initial homogeneous region that turns into a heterogeneous region where the PS precipitates and it is separated from the initial mixture. To determine the homogeneous and heterogeneous regions, experiments were performed at mild working conditions in the temperature and pressure ranges from 25 to 40 ºC and 50 to 150 bar. Higher temperatures were not used to decrease the energetic consume and not exceed the glass transition temperature of the polymer (dramatically decrease by the presence of CO2) (Chapter 4) [16-18]. The initial concentration of PS in Limonene has been considered in the range from 0.05 to 0.80 g/ml since at fixed temperature and pressure, the composition of the phases depends on the initial concentration of the mixture. The presence of solute in a mixture solvent/antisolvent may change the phase behaviour, particularly, in the case of polymers. There are different considerations about the relevance of the solute presence in the 122 Chapter 5 morphology of the precipitated particles or foams, by this reason we decided to include the polymer initial concentration as a variable of study. To test the reproducibility of the experiments, we checked the mass balance of each component, in the two regions, which was better than 2.08 % in vapour phase. The experimental data have been illustrated in a two-dimensional diagram by using an isothermal pressure-composition prism with a triangle as a base [19] (Figures 5.1, 5.2 and 5.3). The binary equilibrium data previously obtained are also included as the limits of the different axis. On analysis of the experimental ternary diagrams (Figure 5.1, 5.2 and 5.3) a twophase region is observed. According to experimental data depicted, it is observed that Polystyrene is not collected in the vapour phase. Samples from vapour phase were analyzed by GPC and any peak belonging to the polymer was detected. At low carbon dioxide concentrations the system is homogeneous when PS concentration in Limonene is lower than around 30% and at higher carbon dioxide concentrations the phase boundary is reached. When PS concentration increases, two phases are made, liquid-liquid phase separation is produced and a polymer rich (and Limonene lean) or lean (and Limonene rich) phases are stood out. When pressure increased, the two one-phase regions slightly shrink; it means that the mutual miscibility of the three components increases as pressure increased, as it was studied with the binary equilibriums (Chapter 4). This fact is better observed when pressure reaches 84.4 bar at 40ºC, after the cited point Limonene is fully miscible with CO2. When temperature increases, two effects are noticeable, by one side, higher temperature improves the volatility of Limonene [20, 21], but on the other side, PS solubility in Limonene is also enhanced [22]. Furthermore, the two-phase region of the Limonene/carbon dioxide system increases, while the solubility of carbon dioxide in PS decreases with increasing temperature [23]. Nevertheless, it was observed that the position of the binodal was not significantly affected by changes in pressure or temperature, in the studied range 50-150 bar and 25-40 ºC. In ternary diagrams the tie lines are also represented, in the two-phases region. They connect the compositions of bottom phase (rich in polymer) and top phase (rich in CO2). As it is observed, these dashed lines present negative slope probably due to the higher interaction between Limonene and PS than with CO2, especially at high initial polymer concentration into the solutions. 123 Figure 5.3. Phase behaviour of the system CO2/Limonene/PS at 40ºC at 50, 75, 100 and 150 bar. Concentrations are given in mass fraction. Figure 5.2 Phase behaviour of the system CO2/Limonene/PS at 30ºC at 50, 75, 100 and 150 bar. Concentrations are given in mass fraction. Figure 5.1. Phase behaviour of the system CO2/Limonene/PS at 25ºC and at 50, 75, 100 and 150 bar. Concentrations are expressed as mass fraction. Equilibrium of Ternary Mixtures 124 Chapter 5 In the ternary representation it is difficult to observe the influence of pressure, temperature and gas content on the phase behaviour at the same time. Next, we have analyzed particularly each variable to better understand their influence in the Limonene solubility in CO2 rich phase. Limonene solubility was subject of study to obtain a polymer similar to the virgin and completely free of solvent. Thus, maximum Limonene solubility in CO2 is the target of this section. 5.2.1. Effect of pressure Figure 5.4 shows the limonene solubility in the vapour phase obtained experimentally in the ternary system CO 2/Limonene/PS. Also the theoretical solubility (represented as lines) is shown from 25 to 40ºC. 120 T: 25ºC 0.05 g/ml 0.10 g/ml 80 0.20 g/ml S (mg/g) 100 60 40 20 0 T: 30ºC 0.05 g/ml 0.10 g/ml 80 0.20 g/ml S (mg/g) 100 60 40 20 0 T: 40ºC 0.05 g/ml 0.10 g/ml 80 0.20 g/ml S (mg/g) 100 60 40 20 0 20 40 60 80 100 120 140 160 180 200 Pressure (bar) Figure 5.4. Effect of pressure on the experimental solubility of Limonene in top phase (markers) Theoretical solubility predicted at 25ºC (a), 30ºC (b) and 40ºC (c). 125 Equilibrium of Ternary Mixtures With increasing pressure of CO2, the solvent power of the molecule to dissolve the terpene increases, and therefore, the concentration of Limonene in the vapour phase becomes larger. Nevertheless, it is observed that solubility increases with pressure, describing an asymptotic tendency that is kept constant at a pressure value around 100 bar. This fact is due to the increase of CO2 density that is directly related to the dissolving power [24], thus, the density of the saturated CO2 phase increases rapidly with pressure, while the density of the liquid phase, saturated with CO 2, although slightly increasing at low and medium pressure, decreases at higher pressure [25]. 5.2.2. Effect of temperature The effect of temperature, at constant pressure, on the experimental (markers) and theoretical (lines) solubility of Limonene in the top phase is shown in Figure 5.5. 60 (a) P: 50 bar 0.05 g PS/ml Lim 20 0.10 g PS/ml Lim 0.20 g PS/ml Lim 15 10 5 0 30 35 20 10 P: 75 bar 0.05 g PS/ml Lim 0.10 g PS/ml Lim 0.20 g PS/ml Lim 25 30 35 40 Temp (ºC) 120 (c) (d) 100 S (mgLim/g CO2) 80 S (mgLim/g CO2) 30 40 Temp (ºC) 100 40 0 25 (b) 50 S (mgLim/g CO2) S (mgLim/g CO2) 25 60 40 P: 100 bar 0.05 g PS/ml Lim 20 0.10 g PS/ml Lim 0.20 g PS/ml Lim 0 25 30 35 Temp (ºC) 40 80 60 40 P: 150 bar 0.05 g PS/ml Lim 20 0.10 g PS/ml Lim 0.20 g PS/ml Lim 0 25 30 35 40 Temp (ºC) Figure 5.5. Effect of temperature on the experimental solubility of Limonene in top phase (markers). Theoretical solubility predicted at 50 (a), 75 (b), 100 (c) and 150 bar (d). According to Figure 5.5, at low pressures, when increasing temperature Limonene solubility increases as it was observed also by Chang and Chen in the binary system [25]. When pressure increases, the critical pressure of the mixture is reached (Limonene/CO2): at 30 ºC, 80 bar and at 40ºC, 85 bar [20] and Limonene is fully miscible with CO2 above this pressure. After reaching the mixture critical point, at 126 Chapter 5 concentration higher than 0.10 g PS/ml Limonene, when temperature increases solubility decreases. It is important to stand out that polymer solubility in Limonene increases when temperature increases as it was shown in previous work [26], which means a decrease in the solubility of Limonene in the CO2 rich phase. This behaviour will be studied in the subsequent section because as it can be observed, the effect of temperature on solubility is quite hazy because it is influenced by several effects. 5.2.3. Effect of concentration Figure 5.6 shows the influence of concentration increase from 0.05 to 0.80 g/ml on the solubility of Limonene in CO2 rich phase. 25 25 ºC 30 ºC 40 ºC 50 S (mgLim/g CO2) 20 S (mgLim/g CO2) 60 (a) P: 50 bar 15 10 5 0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Conc (g PS/ml Lim) 100 40 30 20 (b) P: 75 bar 25 ºC 30 ºC 40 ºC 10 0 0.00 0.05 0.10 0.15 60 40 20 0 0.0 25 ºC 30 ºC 40 ºC 100 S (mgLim/g CO2) S (mgLim/g CO2) 80 0.25 (d) P: 150bar (c) P: 100 bar 25 ºC 30 ºC 40 ºC 0.20 Conc (g PS/ml Lim) 120 80 60 40 20 0.1 0.2 0.3 0.4 0.5 Conc (g PS/ml Lim) 0.6 0.7 0.8 0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Conc (g PS/ml Lim) Figure 5.6. Effect of initial concentration of PS on the experimental solubility of Limonene (markers) at (a) 50 bar; (b) 75 bar; (c) 100 bar; (d) 150 bar. When the concentration of Polystyrene in the solution increases Limonene is embedded in the polymeric matrix making it more difficult to solve in CO 2, independently of the working pressure. The most pronounced negative slope is observed at 100 bar and at high temperature (40ºC) because at those conditions, CO 2 is saturated of Limonene [20], but when initial PS concentration increases, Limonene is more akin to the polymer [26]. In general, Limonene solubility in vapour phase decreases when the initial concentration increases. 127 Equilibrium of Ternary Mixtures Based on these results, the initial concentration of polymer in Limonene is crucial to carry out the separation process by high pressure CO 2. Because contrary to preliminary belief, high concentrations of PS in Limonene do not assist the global recycling process, since values over 0.40 g PS/ml Limonene decreases dramatically the solvent solubility in CO2, at the studied working pressure and temperature. 5.2.4. Sorption of CO2 into PS/Limonene solutions Phase separation produced during CO2 sorption in solutions of PS is crucial in generation of PS porous matrices or foams. While in the unswollen state the diffusion coefficient for large molecules is negligible, it increases by orders of magnitude in the swollen state. The CO2 sorption into PS/Limonene solutions was determined in a range of pressure from 50 to 150 bar and concentration from 0.05 to 0.20 g PS/ml Limonene at two different temperatures, 30 and 40ºC. Figure 5.7 shows data of CO2 sorption obtained experimentally in solutions of PS/Limonene together with the obtained Shieh and Liu in the binary system CO2/PS [27]. According to Figure 5.7, CO2 sorption in PS/Limonene solutions decreases when pressure increases reaching PS sorption typical values over 80 bar approximately, which can be explained if the critical pressure of Limonene solutions is considered [28]. Above this pressure, the terpene oil is fully miscible with CO 2 and the gas molecule could only be retained in the polymer rich phase which is mainly composed by PS and the solvent not solubilised into CO2. As Limonene solubility does not increase once critical mixture conditions are reached, the amount of CO 2 soaked are kept constant. 0.45 T:30ºC; C0: 0.05 gPS/ml Lim T:30ºC; C0: 0.10 gPS/ml Lim 0.40 T:30ºC; C0: 0.20 gPS/ml Lim Sorption (g CO 2/g PS) 0.35 T:40ºC; C0: 0.05 gPS/ml Lim 0.30 T:40ºC; C0: 0.10 gPS/ml Lim 0.25 T:32ºC. Shieh and Liu, 2003 [27] T:40ºC; C0: 0.20 gPS/ml Lim 0.20 0.15 0.10 0.05 0.00 0 25 50 75 100 125 150 175 200 Pressure (bar) Figure 5.7. Sorption data for Polystyrene/Limonene solutions exposed to carbon dioxide. Experimental data from this work at 30ºC and 40ºC are shown as markers (Average Standard Deviation 0.42 mg) and data from Shieh and Liu [27] at 32ºC are shown as a solid line. 128 Chapter 5 Experimental sorption data of CO2 into PS obtained from Shieh and Liu in PS [34] are shown in Figure 5.7. According to the experimental data shown in the literature, when pressure increases the sorption of CO2 increases, reaching a constant value close to its critical point [35]. Comparing our experimental data and the summarized from the literature, it is observed that higher values of sorption are obtained in the case of PS/Limonene solutions, which suggests that the sorption of CO2 is not exclusively stayed into the polymer but also into the Limonene. Nevertheless, when the critical pressure of the mixture is reached, Limonene is fully solubilised in CO2 and it goes into the vapour phase, obtaining values of sorption similar to the described in the literature because the polymeric rich phase was mainly composed of pure PS. 5.2.5. Distribution coefficient Considering the interest of a separation process design for a ternary mixture and taking into account that the CO2 used as supercritical solvent is removed from the system after the separation, a global process can be designed. With respect to a separation process, the partition coefficients (Ki) are more relevant than the solubility. Ki is a key thermodynamic parameter, which is defined as the ratio of the mole fraction of the component in the top phase (yi) to that in the bottom phase (xi) and it is expressed by the following equation. The distribution coefficient of Limonene as a function of pressure, temperature and concentration is depicted in Figure 5.8. Limonene distribution coefficient increases when temperature and pressure increase, having similar shapes to the solubility curves (Figures 5.4 and 5.5.). The effect of concentration can also be observed, when concentration increases from 0.05 to 0.20 g PS/ml Limonene, KLim decreases. According to the separation process a maximum KLim is desired. It can be achieved at pressure close to 100 bar (the plateau is reached), higher temperature and low polymer concentration. 129 0 80 0.015 0.012 25ºC; 0.05 g PS/ml Lim 25ºC; 0.10 g PS/ml Lim 25ºC; 0.20 g PS/ml Lim KLim 0.009 0.006 0.003 90 100 110 120 130 140 150 160 0.000 Pressure (bar) 0.005 KLim 0.004 0.003 0.002 30ºC; 0.05 g PS/ml Lim 30ºC; 0.10 g PS/ml Lim 30ºC; 0.20 g PS/ml Lim 0.001 90 100 110 120 130 140 150 0.000 0.015 160 Pressure (bar) 0.012 KLim 80 Equilibrium of Ternary Mixtures 40ºC; 0.05 g PS/ml Lim 40ºC; 0.10 g PS/ml Lim 40ºC; 0.20 g PS/ml Lim 0.009 0.006 0.003 0.000 40 60 80 100 120 140 160 Pressure (bar) 90 100 110 120 130 Pressure (bar) 140 150 160 Figure 5.8. Distribution coefficient of Limonene at 25, 30 and 40ºC. According to the results described along this section, it can be concluded that it is possible to separate polymer and Limonene rich phases from an initial solution PS/Limonene, using CO2 as antisolvent. The most suitable operational conditions to achieve the Limonene recovery, and subsequently dried Polystyrene recovery, were determined according to experimental results of distribution coefficients. As it was showed, it is possible to solubilise the maximum amount of Limonene at higher pressure and temperature and lower concentrations. 130 Chapter 5 5.3. The ternary system CO2/p-Cymene/PS In the antisolvent techniques the supercritical fluid is not able to dissolve the solid (PS in this case) but it is completely miscible with the liquid. The precipitation process will take place at those conditions where solvent and antisolvent are miscible over the whole composition range while the polymer and the antisolvent are totally immiscible. The development of this section has been divided in four parts: the definition of the ternary diagrams to limit the regions, the study of the solubility of the terpene in the CO2 rich phase, the solubility of the CO2 into the polymeric solution and finally the selection of the most suitable conditions to perform the precipitation of PS from its solution in p-Cymene using CO2. Figure 5.9 shows the vapour-liquid equilibrium composition of the binary system CO2/p-Cymene. According to the Figure, temperatures between 30 and 40 ºC and pressures up to 120 bar would be suitable to solubilise fully the p-Cymene and therefore to precipitate the Polystyrene. According to the VLE data of p-Cymene/CO2 (Figure 5.9), all the experiments were carried out at mild working conditions in the temperature and pressure ranges from 25 to 40 ºC and 50 to 150 bar. The working conditions were selected attending to the experimental solubility of p-Cymene in CO2 in order to achieve the fully solubilisation of p-Cymene in the vapour phase, while the energy consumption was moderate and the glass transition temperature of the polymer was not exceeded (dramatically decrease by the presence of CO 2) [16-18]. The concentration was ranged from 0.05 to 0.80 g PS/ml p-Cymene because it would allow tuning the morphology of the PS during the precipitation hence its importance on the selection of working conditions. At the equilibrium conditions of this work there were two phases: vapour phase enriched in CO2 and liquid phase enriched in PS. It was assumed and checked experimentally that the polymer kept in the liquid phase [9, 29]. To test the reproducibility of the experiments, we checked the mass balance of each component, in the two regions, which was better than 1.14 % in vapour phase and 3.86 % in the liquid phase. 5.3.1. Ternary Diagrams From the binary systems described in the literature (CO 2/p-Cymene [30], pCymene/PS [22] and PS/CO2 [31]), only the limits of the ternary phase diagram could be generated. But the overall knowledge of the system CO 2/p-Cymene/PS must be experimentally obtained since the presence of a solute in a mixture solvent/antisolvent may change the phase behaviour, particularly if the solute is a polymer. 131 Equilibrium of Ternary Mixtures 120 (a) 110 T: 30ºC T: 50ºC T: 70ºC 100 90 Pressure (bar) 80 70 60 50 40 30 20 10 0 0.0 0.1 0.2 0.3 0.4 0.5 Solubility of CO 2 (g CO 2/g Total) 3 0.7 2 (b) Solubility of p-Cymene (kg/m ) 0.6 0.8 0.9 0.99 1.00 x/y CO 120 100 80 60 40 20 0 200 400 600 800 3 Density of CO2 (kg/m ) 1000 3 (c) 2 1 0 0 200 400 600 800 1000 0.8 0.6 0.4 0.2 25 50 75 100 125 150 175 3 Density of CO2 (kg/m ) Figure 5.9. (a) Equilibrium phase composition diagram for the CO 2/p-Cymene binary system at 30, 50 and 70 ºC. (b) Solubility of p-Cymene in the vapour rich phase as a function of density. (c) Effect of density on the solubility of CO 2 into the liquid phase. Homogeneous and heterogeneous regions should be delimited as initial step for the design of a process to recover a polymer from a solution using CO 2 as antisolvent. Next, the selection of the operating conditions should be performed according to the initial homogeneity of the solution which is turned into a heterogeneous mixture in order to precipitate the polymer and separate the solvent. The experimental data have been depicted in a diagram by using an isothermal pressure-composition prism with a triangle as a base [19]. From the analysis of the experimental ternary diagrams (Figure 5.10) the two-phase region was observed (shaded area). At atmospheric pressure, the system is homogeneous till reaches the critical concentration of the polymer ( 29% of PS in weight) [22]. But when CO2 is added to the system, the concentration of PS necessary to induce the phase separation decreases, due to the antisolvent effect of CO 2; the phase boundary is reached and the phase separation occurs. When pressure increases, the solubility of p-Cymene increases and the heterogeneous area was expected to shrink because the mutual miscibility of the three compounds was enhanced. An increase of temperature 132 Chapter 5 entails two noticeable effects, the improvement of p-Cymene solubility in CO2 [30] and the enhancement of PS solubility in p-Cymene [22]. Also, the two-phase region of the p-Cymene/carbon dioxide system increased, while the solubility of carbon dioxide in PS decreased when temperature increased [23]. The increase of PS concentration does not show any crucial shift on the position of the phase boundary. The tie lines shown in Figure 5.10 represent the composition in the bottom phase (polymer rich phase) and the top phase (CO 2 rich phase). The most relevant effect over the position of the tie lines was caused by the PS concentration in the initial solution. The increase in the polymer concentration is the responsible of the increase in the negative slope of the cited lines, probably due to the higher interaction between p-Cymene and PS than with CO2. When pressure rises, the solubility of pCymene in the vapour rich phase increases and the slope of the tie-lines become more pronounced. The position of the binodal line was not significantly influence by shifts in pressure or temperature in the range of the studied conditions. According to Fig. 5.8, the effect of pressure and temperature was not evident from the ternary diagrams since the position of the binodal was not significantly affected by changes in the cited variables. By these reasons, the analysis of the pressure, temperature and concentration on the solubility of p-Cymene in the vapour rich phase and the CO 2 in the polymer rich phase was studied. The pressure and temperature effects was gathered and expressed through the density of CO 2 because it compiles the solvent power of the mixture. 5.3.2. Influence of density and concentration on p-Cymene solubility in the vapour phase The influence of the CO2 density on the solubility of p-Cymene in the CO2 rich phase was studied in the range of pressure from 50 to 150 bar and temperature from 25 to 40 ºC. Figure 5.11 shows the effect of density on the solubility of p-Cymene in the vapour phase at three different concentration of polymer. According to Figure 5.11, when the density of CO 2 increases an increase on pCymene solubility is observed. The most pronounced effect is noticed from 700 kg/m3 belonging to the critical point of the CO 2. This fact demonstrates the improved solvent power of the CO2 above its critical point, which enhances the dissolution of the terpene in the vapour phase. Nevertheless, when temperature increases a density decrease is produced and consequently a decrease on p-Cymene solubility. 133 Equilibrium of Ternary Mixtures Figure 5.10. Phase behaviour of the system CO2/p-Cymene/PS at (a) 25 ºC, (b) 30 ºC and (c) 40 ºC at constant pressure: 50, 100 and 150 bar. Concentrations are given in mass fraction. 134 Chapter 5 150 (a) Binary system CO2/p-Cymene C0:0.05 g PS/ml Cym 120 90 60 150 (b) 3 Solubility of p-Cymene (kg/m ) 30 Binary system CO2/p-Cymene C0:0.10 g PS/ml Cym 120 90 60 30 150 (c) Binary system CO2/p-Cymene C0:0.20 g PS/ml Cym 120 90 60 30 0 0 150 300 450 600 750 900 3 Density of CO2 (kg/m ) Figure 5.11. Influence of CO2 density on the solubility of p-Cymene (kg p-Cymene/m3 CO2) in the vapour rich phase at three different concentration (a) ■ 0.05 gPS/ml Cymene; (b) ● 0.10 gPS/ml Cymene; (c) ▲ 0.20 gPS/ml Cymene. Dashed lines represent the solubility of pCymene in the binary system CO 2/p-Cymene. Also, it is showed that the presence of PS prevents the maximum solubility of pCymene in the vapour phase from being reached. The PS effect is observed by comparison between the solubility of p-Cymene in the binary system CO2/p-Cymene (dashed lines) and in the ternary system CO2/p-Cymene/PS (markers and continuous lines) in Figure 5.11. When there is not any PS in the solution, the highest values of solubility in the binary system are around 150 kg p-Cymene/m3 CO2 (: 900 kg/m3). Nevertheless, when the concentration of polymer is 0.05 g PS/ml p-Cymene, the solubility of p-Cymene in the CO2 rich phase at the same working conditions is around 90 kg p-Cymene/m3 CO2. With the aim of thoroughly study the effect of PS concentration of the initial solution on the solubility of p-Cymene a wider range of concentration was investigated (Figure 5.12). 135 Equilibrium of Ternary Mixtures 100 45 3 0.5 0.0 0.00 2.0 0.05 0.10 0.15 0.20 0.25 Concentration (g PS/ml Cym) 30 15 0 0.00 60 0.05 0.20 0.25 0.5 0.0 0.00 0.05 0.10 0.15 0.20 0.25 Concentration (g PS/ml Cym) 30 15 0 0.00 20 0.5 0.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Concentration (g PS/ml Cym) 0 0.00 80 0.05 0.10 0.15 0.20 0.25 Concentration (g PS/ml Cym) 0.05 0.10 0.15 0.20 (h) P: 150 bar; T: 30 ºC 60 40 20 0 0.00 30 25 0.05 0.10 0.15 0.20 0.25 Concentration (g PS/ml Cym) (i) P: 150 bar; T: 40 ºC 10 20 5 0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Concentration (g PS/ml Cym) 15 10 5 0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Concentration (g PS/ml Cym) Figure 5.12. Effect of concentration of PS (abscissa axis) on the solubility of p-Cymene (kg pCymene/m3 CO2) in the vapour rich phase at (a) P: 50 bar; T: 25 ºC; (b) P: 50 bar; T: 30 ºC; (c) P: 50 bar; T: 40 ºC; (d) P: 100 bar; T:25 ºC; (e) P: 100 bar; T: 30 ºC; (f) P: 100 bar; T: 40 ºC; (g) P: 150 bar; T: 25 ºC; (h) P: 150 bar; T: 30 ºC; (i) P: 150 bar; T: 40 ºC. As it is observed, an increase in the PS concentration entailed a decrease in the solubility of p-Cymene in the CO2 rich phase. When the concentration of PS in the solution increases, p-Cymene is swelling the polymeric solution and the dissolution of the terpene in CO2 is not enhanced. In all cases, at low PS concentration ( 0.1 g PS/ml p-Cymene) there is a decrease of solubility regarding the concentration, but at higher concentration the solubility decrease is softer. Thus, the decrease of pCymene solubility in CO2 as a consequence of the polymer in the initial solution has been confirmed. Polystyrene precipitation process would be preferably achieved at low concentration in order to recover a dried polymer since at those conditions the solvent could be fully soluble in CO2. The nonlinear dependence of density and concentration on the solubility of pCymene was studied. Chrastil’s equation [23] was selected since it has been widely used to describe the phase behaviour of carbon dioxide and terpenes. But according 136 0.25 Concentration (g PS/ml Cym) 3 15 3 3 1.0 25 (f) P: 100 bar; T: 40 ºC S (kg Cym/m CO2) (c) P: 50 bar; T: 40 ºC 1.5 50 3 45 3 1.0 2.0 S (kg Cym/m CO2) 0.15 (e) P: 100 bar; T: 30 ºC S (kg Cym/m CO2) 3 S (kg Cym/m CO2) (b) P: 50 bar; T: 30 ºC 1.5 0.10 Concentration (g PS/ml Cym) S (kg Cym/m CO2) 1.0 (g) P: 150 bar; T: 25 ºC 75 3 S (kg Cym/m CO2) 1.5 3 S (kg Cym/m CO2) (d) P: 100 bar; T: 25 ºC S (kg Cym/m CO2) 60 (a) P: 50 bar; T: 25 ºC S (kg Cym/m CO2) 2.0 Chapter 5 to the results shown in Figure 5.11 and 5.12, a modification of the cited equation was proposed to include the effect of polymer concentration in the initial solution. [5.1] where S is the solubility of p-Cymene in CO2 rich phase in [kg/m3], C1 is a constant dependent on the molecular weights of the solute and solvent and on the association constant, C2 is a constant dependent on the total heat of vaporization, T is the temperature in [K], k is the association number of molecules in the solvate-complex, is the CO2 density in [kg/m3], C3 is the new fitting constant and CPS is the concentration of PS in the initial solution in [g PS/ml p-Cymene]. The results of the fitting and deviations are shown in Table 5.1. Table 5.1. Estimated parameters obtained from the correlation of eq. [5.1] for the solubility of p-Cymene in vapour phase as a function of CO 2 density, temperature and concentration of PS in the initial solution. C1 C2 C3 k R2 -43.26 ± 2.06 1460.15±614.02 -0.51±0.04 6.09±0.33 0.93 The correlation fits the experimental data very closely and presents an acceptable value of correlation in a range of pressure from 50 to 150 bar, temperature from 25 to 40 ºC and concentration from 0.05 to 0.80 g PS/ml p-Cymene. The opposed effects of density and concentration on the solubility of p-Cymene in CO2 from the ternary system are observed through the opposite sign of k and C3, where the positive influence of density prevails over the negative effect of PS concentration. In Figure 5.13 the comparison between solubility of p-Cymene correlated and experimental values is observed. 80 3 Solubilitycorrel (kg p-Cymene/m CO2) 100 60 40 20 0 0 10 20 30 40 50 60 70 80 90 100 3 Solubilityexp (kg p-Cymene/m CO2) Figure 5.13. Comparison between experimental and correlated solubility with the modified Chrastil’s equation of p-Cymene in CO2. 137 Equilibrium of Ternary Mixtures The modification of the Chrastil’s semi-empirical equation including the effect of the polymer concentration fits suitably the experimental results for the ternary system CO2/p-Cymene/PS. 5.3.3. Influence of density and concentration on CO2 solubility in the liquid phase The CO2 absorbed into the polymeric solution would determine the feasibility to carry out the antisolvent process and the possible modification of the PS structure. The effect of density on the CO2 sorption in the liquid phase in a range of pressure from 50 to 150 bar, temperature from 25 to 40 ºC and at three different values of concentration (0.05-0.10-0.20 g PS/ml Cymene) is shown in Figure 5.14. 0.20 (a) Binary system CO2/PS C0:0.05 g PS/ml Cym 0.15 Sorption of CO 2 (g CO 2/g Solution) 0.10 0.05 0.00 (b) 0.12 Binary system CO2/PS C0:0.10 g PS/ml Cym 0.09 0.06 0.03 0.00 (c) 0.08 Binary system CO2/PS C0:0.20 g PS/ml Cym 0.06 0.04 0.02 0.00 0 150 300 450 600 750 900 3 Density of CO2 (kg/m ) Figure 5.14. Influence of CO2 density on the solubility of CO2 (g CO2/ g Solution) in the polymeric rich phase in a range of concentration from (a) ■ 0.05 gPS/ml Cymene; (b) ● 0.10 gPS/ml Cymene; (c) ▲ 0.20 gPS/ml Cymene. 138 Chapter 5 The influence of density on CO2 sorption inside the solution shows that when density increases, the sorption of CO2 in the polymeric rich phase (PS/p-Cymene) decreases. Independently of the PS concentration in the initial solution, the CO 2 sorption reaches a constant value at density above 600 kg/m3. Figure 5.14 shows that the sorption of CO2 decreased up to 300-400 kg/m3 where it is kept practically constant around 0.02 g CO2/g of solution. According to the literature, the sorption of CO2 in PS at 100 bar and 32 ºC is close to 0.05 g CO2/ g PS which agrees with the data obtained in this work [24]. Nevertheless, the highest value of CO2 sorption in the ternary system is reached at C: 0.05 g PS/ ml p-Cymene and : 100 kg/m3 ( 0.2 g CO2/g Solution). Similar values are observed in the binary system CO2/p-Cymene shown in Figure 5.9 c, which could be explained by the low concentration of polymer into the solution. The comparison among the sorption of CO2 in the binary systems (CO2/p-Cymene and CO2/PS) with the values of sorption obtained in this work in the case of the ternary system (CO2/pCymene/PS) at concentration 0.05 g PS/ ml p-Cymene are shown in Figure 5.15. 1.0 Binary system CO2/p-Cymene Binary system CO2/PS Ternary system CO2/p-Cymene/PS Sorption of CO 2 (g CO2/g Total) 0.8 0.6 0.4 0.2 0.0 0 150 300 450 600 3 750 900 Density of CO2 (kg/m ) Figure 5.15. Comparison between the solubility of CO2 into p-Cymene (■), PS (●) and in the solution PS/p-Cymene (*) as a function of density. At low values of density, the sorption of CO 2 in the polymeric solution is similar to the values of solubility in the binary system CO2/p-Cymene; but when density increases, the sorption of CO2 decreases reaching similar values to the solubility of CO2 in Polystyrene. According to Figure 5.15, the behaviour of the solubility of CO2 in the polymeric solutions could be explained attending to the presence of p-Cymene in the mixture which could modify the sorption of CO 2. It can be inferred that CO2 sorption could be distributed between the solvent and the polymer following the next expression. 139 Equilibrium of Ternary Mixtures [5.2] where wCO2sol is the total amount of CO2 retained in the solution, wCO2PS and wCO2Cym represents the CO2 retained in PS and p-Cymene, respectively, and is a fitting parameter. wCO2sol was obtained experimentally; the CO2 sorption in PS (wCO2PS) was calculated following Sanchez-Lacombe equation of state and data were compared with the literature [25-28]. The CO2 absorbed in p-Cymene (wCO2Cym) was obtained according to Peng-Robinson equation of state and they were also checked with the literature [19] and our own data. is a fitting variable depending on the pressure, temperature and initial concentration of the solution that was obtained from eq. [5.2]. The influence of pressure and temperature (expressed as density) on is shown in Figure 5.16, where decreased when CO2 density increased and reached a constant value near zero when p-Cymene is fully soluble in CO2 and it is only retained inside the PS. 0.30 (a) C0:0.05 g PS/ml Cym 0.25 0.20 0.15 0.10 0.05 0.00 (b) C0:0.1 g PS/ml Cym (c) C0:0.20 g PS/ml Cym 0.12 0.09 0.06 0.03 0.00 0.08 0.06 0.04 0.02 0.00 0 150 300 450 600 750 900 3 Density of CO2 (kg/m ) Figure 5.16. Influence of density and initial concentration of PS/p-Cymene solutions on the parameter at (a) ■ 0.05 gPS/ml Cymene; (b) ● 0.10 gPS/ml Cymene; (c) ▲ 0.20 gPS/ml Cymene. 140 Chapter 5 An increase on CO2 density promotes an increase of the CO2 absorbed in p-Cymene that reaches the saturation around 400 kg/m3 corresponding with the critical point of the mixture ( 84 bar and 40ºC) [9]. This behaviour is expected since the amount of available p-Cymene for the CO2 to be absorbed in is lower when the PS concentration increases. The solvent is the responsible to increase the sorption of CO2 inside the solution but when the critical point of the mixture is reached, the sorption is exclusively due to the polymer. Thus, the decrease of CO2 sorption in the PS/p-Cymene mixtures at low pressures is due to the solubilisation of p-Cymene in the vapour rich phase which reduced the capability of sorption of the solutions since the solvent is volatilised. The effect of temperature on the CO2 sorption shows that an increase of temperature entails a decrease of density and consequently an increase on CO 2 solubility inside the PS/pCymene solutions. The influence of the concentration of PS in the p-Cymene solution on the sorption of CO2 into the polymeric rich phase was studied from 25 to 40º C, at three different pressures (50-100-150 bar) and in a wider range of concentration from 0.05-0.80 g PS/ml p-Cymene (Figure 5.17). According to Figure 5.17, the sorption of CO2 into the polymeric solutions decreases when pressure increases and temperature decreases, it means, when the density of CO2 increases. At constant temperature, when pressure increases, the solvent power of CO2 is enhanced and the amount of p-Cymene in the solution decreases since it is in vapour phase and consequently, the concentration of CO 2 into the polymer liquid phase decreases. The effect of density and concentration on the sorption of CO 2 in the polymer rich phase could be fitted following a modification of the Dual-Mode sorption model [29, 30] (Chapter 4). It is typically used to describe the sorption of gases in polymers but in this work the innovation proposes the replacement of pressure by density and the addition of a third term to represent the polymer concentration: [5.3] where S is the solubility of CO2 in the polymeric rich phase in [g CO2/g solution], kD is the Henry’s law coefficient, b represents the hole affinity parameter, which is a measure of the affinity between the solute molecules and the Langmuir sites, C’ H is the capacity parameter, is the density replacing the traditional value of pressure shown in the model to include the temperature effect in [kg/m3], C’P and a are the new variables introduced to include the concentration of polymer and C is the concentration of PS in the initial solution in [g PS/ml p-Cymene]. 141 Equilibrium of Ternary Mixtures S (g CO 2/g Solution) S (g CO 2/g Solution) 0.06 0.03 0.04 0.02 0.02 0.00 0.00 0.05 0.10 0.15 0.20 0.01 0.00 0.25 0.00 Concentration (g PS/ml Cym) (b) P: 50 bar; T: 30 ºC 0.05 0.10 0.15 0.20 0.25 0.10 Concentration (g PS/ml Cym) (e) P: 100 bar; T: 30 ºC 0.08 0.15 0.20 0.05 0.10 0.15 0.20 0.25 Concentration (g PS/ml Cym) (c) P: 50 bar; T: 40 ºC 0.05 0.10 0.15 0.20 0.25 Concentration (g PS/ml Cym) (f) P: 100 bar; T: 40 ºC 0.08 0.15 0.05 0.05 0.10 0.15 0.20 0.25 Concentration (g PS/ml Cym) (i) P: 150 bar; T: 40 ºC 0.01 0.00 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Concentration (g PS/ml Cym) 0.00 0.00 0.02 0.02 0.00 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.25 0.03 0.04 0.05 0.20 0.04 0.06 0.10 0.15 0.01 0.00 0.00 0.10 0.10 0.02 0.02 0.00 0.00 0.05 Concentration (g PS/ml Cym) (h) P: 150 bar; T: 30 ºC 0.03 0.04 0.05 0.05 0.04 0.06 0.10 0.00 0.00 S (g CO 2/g Solution) 0.05 S (g CO 2/g Solution) 0.10 (g) P: 150 bar; T: 25 ºC 0.04 S (g CO 2/g Solution) S (g CO 2/g Solution) 0.05 (d) P: 100 bar; T: 25 ºC 0.08 0.15 0.20 S (g CO 2/g Solution) 0.10 (a) P: 50 bar; T: 25 ºC S (g CO 2/g Solution) S (g CO 2/g Solution) 0.20 0.00 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Concentration (g PS/ml Cym) Concentration (g PS/ml Cym) Figure 5.17. Influence of CO2 density on the solubility of CO2 (g CO2/ g Solution) in the polymeric rich phase from an initial PS/p-Cymene solution at constant pressure and temperature (a) P: 50 bar; T: 25 ºC; (b) P: 50 bar; T: 30 ºC; (c) P: 50 bar; T: 40 ºC; (d) P: 100 bar; T:25 ºC; (e) P: 100 bar; T: 30 ºC; (f) P: 100 bar; T: 40 ºC; (g) P: 150 bar; T: 25 ºC; (h) P: 150 bar; T: 30 ºC; (i)c P: 150 bar; T: 40 ºC. The results of the fitting are shown in Table 5.2. The value of correlation was acceptable especially at high density since the model is recommended for glassy polymers which exhibit lower values of CO2 sorption. Table 5.2. Estimated parameters obtained from the correlation of eq. [5.3] for the sorption of CO2 in the polymeric rich phase as a function of CO2 density, and concentration of PS in the initial solution. kD (m3/kg) C’H b (m3/kg) C’P a (ml/g) R2 -9.02·10-5 ± 2.38·10-5 0.19±0.07 -0.19±0.04 -0.14±0.06 22.46±21.10 0.78 Figure 5.18 shows the comparison between the experimental sorption of CO 2 into the polymeric rich phase and the correlation according to equation [5.3]. 142 Chapter 5 Sorptioncorrel (g CO2/g Total) 0.100 0.075 0.050 0.025 0.000 0.000 0.025 0.050 0.075 0.100 Sorptionexp (g CO2/g Total) Figure 5.18. Comparison between experimental and correlated solubility with the modified dual-modal equation of CO2 in the polymeric solution PS/p-Cymene. The modification of the Dual-Mode sorption including the influence of the concentration in the initial solution fits pretty well the experimental results. 5.3.4. Selection of the operating conditions Considering the interest of a precipitation process for the ternary mixture CO2/pCymene/PS and taking into account that the CO 2 used as supercritical solvent is removed from the system after the separation, the optimum conditions could be selected. With respect to a separation process, the partition coefficients (Ki) are more relevant than the solubility. Ki is a key thermodynamic parameter, which is defined as the ratio of the mole fraction of the component i (CO 2 or p-Cymene) in the top phase (yi) to that in the bottom phase (xi) and it is expressed by the equation. [5.4] The distribution coefficients of CO2 and p-Cymene as a function of pressure, temperature and concentration were depicted in Figure 5.19. CO2 and p-Cymene distribution coefficient increases when pressure increases, but decreases when temperature increases. The effect of concentration shows a negative influence in the case of p-Cymene distribution coefficients, especially observed at 40 ºC but a positive effect on the partition coefficient of CO 2. According to the experimental data, when PS concentration increases the solubility of p-Cymene and CO2 in the vapour or liquid phase, respectively, decreases. The maximum solubility of p-Cymene in the vapour phase is reached at the highest values of density, while 143 Equilibrium of Ternary Mixtures the highest values of CO2 sorption in the polymeric solutions are achieved when the density of CO2 is low, showing opposite trends. Thus, the most suitable conditions to perform the antisolvent process would be selected on the basis of the maximum solubility of the solvent in the vapour phase since the solubility of CO 2 in the solution could be considered constant and equal to the sorption in PS. 20 0.04 25 ºC 25 ºC C0: 0.05 g PS/ml p-Cymene 16 C0: 0.05 g PS/ml p-Cymene C0: 0.10 g PS/ml p-Cymene 0.03 kp-Cymene 2 kCO C0: 0.10 g PS/ml p-Cymene C0: 0.20 g PS/ml p-Cymene 12 8 C0: 0.20 g PS/ml p-Cymene 0.02 0.01 4 20 0.030 30 ºC 30 ºC C0: 0.05 g PS/ml p-Cymene 16 C0: 0.10 g PS/ml p-Cymene C0: 0.10 g PS/ml p-Cymene 0.020 C0: 0.20 g PS/ml p-Cymene 2 kp-Cymene 12 kCO C0: 0.05 g PS/ml p-Cymene 0.025 8 4 C0: 0.20 g PS/ml p-Cymene 0.015 0.010 0.005 15 0.015 40 ºC 40 ºC C0: 0.05 g PS/ml p-Cymene 12 C0: 0.05 g PS/ml p-Cymene 0.012 C0: 0.10 g PS/ml p-Cymene C0: 0.10 g PS/ml p-Cymene C0: 0.20 g PS/ml p-Cymene kp-Cymene kCO C0: 0.20 g PS/ml p-Cymene 0.009 2 9 6 3 0.006 0.003 0 0.000 0 25 50 75 100 125 Pressure (bar) 150 175 0 25 50 75 100 125 150 175 Pressure (bar) Figure 5.18. Distribution coefficient of CO2 (left) and p-Cymene (right) as a function of temperature (25, 30 and 40ºC) and concentration (■) 0.05 g PS/ ml p-Cymene; (○) 0.10 g PS/ ml p-Cymene; (▲) 0.20 g PS/ ml p-Cymene. New operating conditions could be tried theoretically following the methodology described in this work for the recovery of PS. Employing the correlation described in equations [5.1] and [5.3] it should be possible to predict the solubility of p-Cymene in CO2 and the sorption of CO2 into the solution in order to optimize the precipitation. The aim of this work is the recycling of PS wastes. By this reason, the maximum amount of polymer should be recovered without traces of solvent. Furthermore, the process should be economically interesting which implies the minimization of pressure and temperature. In accordance with these premises, a set 144 Chapter 5 of pressure, temperature and concentration values were studied theoretically to define the optimum conditions where the precipitation could be performed. The selected pressure to study the solubility of the system was set at 75 bar while the temperature and concentration was varied from 25 to 40 ºC and from 0.01 to 0.60 g PS/ml p-Cymene, respectively. The ternary diagrams could be predicted and are depicted in Figure 5.20 where the heterogeneous region shows the composition in the polymer and CO2 rich phases. Figure 5.20. Phase diagrams of the ternary mixture CO 2/p-Cymene/PS at 75 bar and 25, 30, 35 and 40 ºC. The partition coefficients were calculated in order to select the most suitable temperature and concentration. As it was explained, the target is the minimization of KCO2 and the maximization of Kp-Cymene while the maximum value of Kp-Cymene should prevail over the minimum of K CO2. Figure 5.21 shows the influence of temperature and concentration on the partition coefficients at 75 bar. 145 Equilibrium of Ternary Mixtures 120 0.040 Pressure: 75 bar Temperature: 25 ºC Temperature: 30 ºC Temperature: 35 ºC Temperature: 40 ºC 0.035 0.030 0.025 kp-Cymene kCO 2 Pressure: 75 bar Temperature: 25 ºC 100 Temperature: 30 ºC Temperature: 35 ºC 80 Temperature: 40 ºC 60 40 0.020 0.015 0.010 20 0 0.0 0.005 0.1 0.2 0.3 0.4 0.5 0.6 Concentration (g PS/ml p-Cymene) 0.7 0.000 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 Concentration (g PS/ml p-Cymene) Figure 5.21. Distribution coefficient of CO2 (left) and p-Cymene (right) as a function of Polystyrene concentration and temperature: () 25 ºC, (○) 30 ºC, (▲) 35 ºC and () 40ºC. According to Figure 5.21, the optimum conditions to perform the precipitation of PS at 75 bar would be 25 ºC (although lower temperature would show better results) and 0.1 g PS/ml p-Cymene to maximize the amount of recovered PS. The selected operating conditions assure the both premises, the maximum solubility of p-Cymene in the vapour phase and of CO2 into the polymeric rich phase. In the light of the results obtained, the separation of Polystyrene from its solutions in terpenes by CO2 at high pressure is feasible due to the differences between the solubility. The most suitable operating conditions to perform the separation process are high pressure, low temperature and moderate concentration since the aim of this work is the recycling of PS wastes into high quality products. 146 Chapter 5 References [1] G. Luna-Bárcenas, S.K. Kanakia, I.C. Sanchez, K.P. Johnston, Semicrystalline microfibrils and hollow fibres by precipitation with a compressed-fluid antisolvent, Polymer, 36 (1995) 3173-3182. [2] S. Mawson, S. Kanakia, K.P. Johnston, Metastable polymer blends by precipitation with a compressed fluid antisolvent, Polymer, 38 (1997) 2957-2967. [3] D.J. Dixon, K.P. Johnston, Formation of microporous polymer fibers and oriented fibrils by precipitation with a compressed fluid antisolvent, Journal of Applied Polymer Science, 50 (1993) 1929-1942. [4] Y. Pérez De Diego, F.E. Wubbolts, G.J. Witkamp, T.W. De Loos, P.J. Jansens, Measurements of the phase behaviour of the system dextran/DMSO/CO2 at high pressures, Journal of Supercritical Fluids, 35 (2005) 1-9. [5] J.L. Owens, K.S. Anseth, T.W. Randolph, Mechanism of microparticle formation in the compressed antisolvent precipitation and photopolymerization (CAPP) process, Langmuir, 19 (2003) 3926-3934. [6] E. Elizondo, A. Córdoba, S. Sala, N. Ventosa, J. Veciana, Preparation of biodegradable poly (methyl vinyl ether-co-maleic anhydride) nanostructured microparticles by precipitation with a compressed antisolvent, Journal of Supercritical Fluids, 53 (2010) 108-114. [7] I. De Marco, E. Reverchon, Nanostructured cellulose acetate filaments produced by supercritical antisolvent precipitation, The Journal of Supercritical Fluids, 55 (2011) 1095-1103. [8] A.I. Cooper, Polymer synthesis and processing using supercritical carbon dioxide, Journal of Materials Chemistry, 10 (2000) 207-234. [9] F. Rindfleisch, T.P. DiNoia, M.A. McHugh, Solubility of polymers and copolymers in supercritical CO2, Journal of Physical Chemistry, 100 (1996) 15581-15587. [10] I. Kikic, M. Lora, A. Bertucco, A Thermodynamic Analysis of Three-Phase Equilibria in Binary and Ternary Systems for Applications in Rapid Expansion of a Supercritical Solution (RESS), Particles from Gas-Saturated Solutions (PGSS), and Supercritical Antisolvent (SAS), Industrial and Engineering Chemistry Research, 36 (1997) 5507-5515. [11] D. Browarzik, M. Kowalewski, Calculation of the stability and of the phase equilibrium in the system polystyrene + cyclohexane + carbon dioxide based on equations of state, Fluid Phase Equilibria, 163 (1999) 43-60. [12] B. Bungert, G. Sadowski, W. Arlt, Supercritical antisolvent fractionation: Measurements in the systems monodisperse and bidisperse polystyrene-cyclohexanecarbon dioxide, Fluid Phase Equilibria, 139 (1997) 349-359. [13] R.A. Krenz, R.A. Heidemann, Modelling the fluid phase behaviour of polydisperse polyethylene blends in hydrocarbons using the modified SanchezLacombe equation of state, Fluid Phase Equilibria, 262 (2007) 217-226. [14] T. Tassaing, P. Lalanne, S. Rey, F. Cansell, M. Besnard, Spectroscopic study of the polystyrene/CO2/ethanol system, Industrial and Engineering Chemistry Research, 39 (2000) 4470-4475. [15] A.P. Mathew, S. Packirisamy, S. Thomas, Studies on the thermal stability of natural rubber/polystyrene interpenetrating polymer networks: Thermogravimetric analysis, Polymer Degradation and Stability, 72 (2001) 423-439. [16] Z. Zhang, Y.P. Handa, An in situ study of plasticization of polymers by highpressure gases, Journal of Polymer Science, Part B: Polymer Physics, 36 (1998) 977982. 147 Equilibrium of Ternary Mixtures [17] C. Tsioptsias, C. Panayiotou, Simultaneous determination of sorption, heat of sorption, diffusion coefficient and glass transition depression in polymer-CO2 systems, Thermochimica Acta, 521 (2011) 98-106. [18] M. Pantoula, C. Panayiotou, Sorption and swelling in glassy polymer/carbon dioxide systems: Part I. Sorption, Journal of Supercritical Fluids, 37 (2006) 254-262. [19] J.M. Prausnitz, R.N. Lichtenthaler, E.G. de Azevedo, Molecular Thermodynamics of Fluid-Phase Equilibria, Pearson Education, 1998. [20] H. Sovová, R.P. Stateva, A.A. Galushko, Essential oils from seeds: Solubility of limonene in supercritical CO2 and how it is affected by fatty oil, Journal of Supercritical Fluids, 20 (2001) 113-129. [21] S.A.B. Vieira de Melo, G.M.N. Costa, A.M.C. Uller, F.L.P. Pessoa, Modeling high-pressure vapor-liquid equilibrium of limonene, linalool and carbon dioxide systems, Journal of Supercritical Fluids, 16 (1999) 107-117. [22] C. Gutiérrez, M.T. García, I. Gracia, A. De Lucas, J.F. Rodríguez, The selective dissolution technique as initial step for polystyrene recycling, Waste and Biomass Valorization, 4 (2013) 29-36. [23] Y. Sato, T. Takikawa, S. Takishima, H. Masuoka, Solubilities and diffusion coefficients of carbon dioxide in poly(vinyl acetate) and polystyrene, The Journal of Supercritical Fluids, 19 (2001) 187-198. [24] M.E. Paulaitis, V.J. Krukonis, R.T. Kurnik, R.C. Reid, Supercritical Fluid Extraction, Reviews in Chemical Engineering, 1 (1983) 179-250. [25] C.-M.J. Chang, C.-C. Chen, High-pressure densities and P-T-x-y diagrams for carbon dioxide+linalool and carbon dioxide+limonene, Fluid Phase Equilibria, 163 (1999) 119-126. [26] C. Gutiérrez, M.T. García, I. Gracia, A. de Lucas, J.F. Rodríguez, Recycling of extruded polystyrene wastes by dissolution and supercritical CO 2 technology, Journal of Material Cycles and Waste Management, 14 (2012) 308-316. [27] Y.T. Shieh, K.H. Liu, The effect of carbonyl group on sorption of CO 2 in glassy polymers, Journal of Supercritical Fluids, 25 (2003) 261-268. [28] C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. De Lucas, M.T. García, Development of a strategy for the foaming of polystyrene dissolutions in scCO2, Journal of Supercritical Fluids, 76 (2013) 126-134. [29] M.L. O'Neill, Q. Cao, M. Fang, K.P. Johnston, S.P. Wilkinson, C.D. Smith, J.L. Kerschner, S.H. Jureller, Solubility of homopolymers and copolymers in carbon dioxide, Industrial and Engineering Chemistry Research, 37 (1998) 3067-3079. [30] Z. Wagner, J. Pavlíček, Vapour-liquid equilibrium in the carbon dioxide- pcymene system at high pressure, Fluid Phase Equilibria, 90 (1993) 135-141. [31] Y. Sato, K. Fujiwara, T. Takikawa, Sumarno, S. Takishima, H. Masuoka, Solubilities and diffusion coefficients of carbon dioxide and nitrogen in polypropylene, high-density polyethylene, and polystyrene under high pressures and temperatures, Fluid Phase Equilibria, 162 (1999) 261-276. 148 Chapter 6 TERNARY SYSTEMS: PROPERTIES Chapter 6 concerns the influence of CO2 on the properties of the solutions is a key factor since the sorption of CO2 plasticises the mixture changing dramatically their physic-chemical and rheological characteristics. According to this, the Chapter has been divided in three different sections: the study of the glass transition temperature, the viscosity and the interfacial tension of the mixtures CO2/Limonene/PS and CO2/p-Cymene/PS. Ternary Systems: Properties 150 Chapter 6 Based on the papers: C. Gutiérrez, M.T. Garcia, S. Curia, S.M. Howdle, J.F. Rodriguez, The effect of CO2 on the viscosity of polystyrene/limonene solutions, Journal of Supercritical Fluids, 88 (2014) 26-37. C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. De Lucas, M.T. García, Development of a strategy for the foaming of polystyrene dissolutions in scCO 2, Journal of Supercritical Fluids, 76 (2013) 126-134. Graphical abstract 151 Ternary Systems: Properties RESUMEN Con el fin de diseñar un proceso para el reciclaje de residuos de Poliestireno mediante su disolución en terpenos y posterior separación mediante CO 2 a alta presión, es necesario estudiar los cambios que éste produce en las propiedades de las disoluciones. El estudio del equilibrio del sistema ternario formado es fundamental para determinar la viabilidad el proceso, sin embargo, también es esencial el estudio de los cambios producidos en la viscosidad, la tensión superficial y la temperatura de transición vítrea de las disoluciones como consecuencia de la plastificación. La viscosidad de las disoluciones de PS/Limoneno en presencia de CO 2 se midió en un viscosímetro de cuarzo. Las condiciones de presión se variaron desde presión atmosférica hasta 50 bar, en un rango de temperatura de 20 a 40 ºC y en un intervalo de concentración de entre 0.025 y 0.40 g PS/ml Limoneno. Se observó que en todos los casos, al aumentar la presión de CO2, la viscosidad de las disoluciones disminuía por el efecto plastificante que ejerce el gas. La determinación de la tensión interfacial se realizó a través del método de la gota colgante en una celda de visión de alta presión bajo condiciones de operación moderadas (presión: 0-90 bar; temperatura: 30-40 ºC). Por último, se estudió el efecto del CO2 sobre la temperatura de transición vítrea del Poliestireno en disolución. A partir de la extrapolación de los datos de viscosidad y tensión interfacial se determinaron los puntos críticos de las disoluciones en presencia de CO2. 152 Chapter 6 ABSTRACT With the aim of designing accurately the global process, the modification of solutions properties in CO2 should be considered. Although the knowledge of the equilibrium is crucial to determine the feasibility of the recycling process, the shifts in the viscosity, the interfacial tension and the glass transition temperature of Polystyrene/Terpene oils solutions should be also studied. The viscosity measurements were performed in a quartz crystal viscometer to determine the viscosity of solutions of Polystyrene in Limonene in the presence of high pressure carbon dioxide. These measurements were determined up to 50 bar in the range of temperature from 20 to 40ºC and in the range of concentration from 0.025 to 0.4 g PS/ml Limonene. The viscosities of the solutions at all loadings were found to decrease with increasing temperature and pressure while the plastising effect of CO2 prevails over the hydrostatic pressure applied by the gas. On the other hand, interfacial tension measurements were carried out by the pendant droplet method at mild working conditions (pressure: 0-90 bar, temperature: 30-40 ºC). Also, the effect of CO2 on the glass transition Polystyrene solutions was studied. Finally, the viscosity and interfacial tension data were used to determine the solvent quality and to quantify the demixing points of the mixtures. 153 Ternary Systems: Properties 154 Chapter 6 6.1. Background of viscosity and interfacial tension of the ternary mixtures CO2/Limonene/PS and CO2/p-Cymene/PS As it has been mentioned, the use of high-pressure fluids for polymer processing and polymer synthesis has grown significantly in recent years [1-3]. Several research groups are focused on the use of supercritical CO 2 as a reaction medium for polymerisation [4-6], as a solvent for fractionation [7, 8], as an antisolvent for encapsulation [9, 10] or as a blowing agent for non-reactive processes [11-13]. For polymer solutions, high-pressure viscosity data are especially important in a variety of applications such as high-pressure synthesis, melt polymer processing and polymer blending [1]. Thus, the viscosity of a polymer coupled with a solvent or plasticizer is a subject of considerable interest for the manufacturing of polymers [14]. A fundamental understanding of how the properties of materials and the processing parameters affect the behaviour of the mixtures is required to optimize the future development of high-quality plastic products. Viscosity also influences non-equilibrium processes, such as those met in the transient stages during phase separation from polymer solutions [15]. Phase separation studies are important in a wide variety of applications, such as spray coating, foaming, spinning (i.e. fibre formation), manufacture of composite materials, membranes and other porous materials. In all these cases, viscosity data of polymer solutions are used to gain fundamental insights into polymer–solvent interactions and to determine if the solvent is a “poor” or a “good” solvent [1] according to their spatial conformation and dissolution ability. In the case of “poor solvents” the polymeric chains stay packed and could form a hard sphere because the solvent can not penetrate easily and dissolve the polymer; nevertheless in the “good solvents” the chains are expanded and the contact between the polymer and the surrounding fluid is maximized. The viscosity of a polymer coil is a property related to the chain dimensions [16]. The degree of polymer-chain expansion depends on several factors but mainly the quality of the solvent which can be altered by changing the temperature, pressure, or even its composition [15]. The pressure dependency of the viscosity of polymer solutions is fundamental to understand the relationship between the polymersolvent interactions in near or supercritical conditions, and the relation between the chain conformation and viscosity. This is because, in addition to temperature, pressure can provide an interesting and useful route to controlling the properties of solutions [1]. The viscosity asymmetry of the components molecules (long chain polymers and smaller solvent) and the interfacial tension of the polymer solutions are mainly responsible for the different morphologies obtained during phase separation: well defined droplets, fibres, membranes, foams etc [15]. Rheological measurements could be used to relate the composition of solutions and the final structure [17]. 155 Ternary Systems: Properties Although polymer solutions in scCO2 are involved in several polymer processing operations, such as fractionation, dyeing, foaming, encapsulation, impregnation, etc. [18-21] there is an important lack of information on their rheological properties. This has been exacerbated by the difficulty in performing experiments at uniform shear rate, at high pressure and under homogeneous conditions. According to the previous section of this Chapter, the knowledge of the equilibrium is very important for the design of the antisolvent processes, but the determination of the thermophysical properties (density, viscosity and interfacial tension) are among the most influential parameters on fluid characteristics. Thus, when a solution is subjected to a quench at the critical polymer concentration, the system enters in the unstable region and phase separation proceeds spontaneously by a mechanism known as spinodal decomposition. For the successful and large scale use of the described recycling process, an intensive investigation of the physico-chemical properties is necessary, especially because of the scarce data shown in the literature. The difference between the viscosity of the solutions and the surrounding medium is one of key contributors to the phase separation since the CO 2 acts as antisolvent of the polymer inducing its precipitation [22]. On the other hand, Interfacial Tension (IFT) is also one of the most important physicochemical parameters used in many polymer engineering processes like foaming, since this property exerts a notable influence on the morphology of the final polymer products. Generally low interfacial tension values are desired to increase the nucleation rate and produce small and uniform cells during the foaming process [23-25]. The nucleation cell is expressed following the Classical homogeneous Nucleation Theory (CNT): Nhom = C· f· exp(−ΔGhom / kB·T) [6.1] where Nhom is the rate of homogeneous nucleation, C is the concentration of nucleation sites, f is the frequency factor of the gas molecules, ΔGhom is the Gibbs free energy for homogeneous nucleation, and kB is Boltzmann’s constant [26, 27]. The energy barrier of cell nucleation can be lowered by decreasing the IFT according the equation below: ΔG = 16 π ϭ3 /3Δp2 [6.2] where ϭ is the surface tension between the polymer phase and nucleating bubble phase, and Δp is the pressure difference across the polymer-gas interface. In summary the classical nucleation theory predicts that the Gibbs free energy will be reduced by the cubic power of the surface tension, and hence the nucleation rate will increase exponentially. As it was shown in Chapter 4, the sorption of supercritical CO2 in polymers results in their swelling and plasticization. Under this condition, carbon dioxide swells the entangled melt polymer, causing an increase in free volume, and consequently, a decrease in chain entanglement. The increase in free volume also increases the 156 Chapter 6 chain mobility, these facts cause changes in their mechanical and physical properties [28, 29], which ultimately can be used to reduce interfacial tension [30, 31]. Furthermore, foaming processes of polymers using CO 2 are generally performed at high temperatures to promote the diffusivity and solubility of the foaming agent in the polymer [32]. In fact, polymer foaming is usually carried out at temperature that exceeds the polymer melting point, to increase mobility in polymer chains and to enhance its deformation [59]. In this work, low temperatures are required for the recycling of PS since the polymer is solubilised in terpenes and they are separated by CO2. The presence of a solvent decreases the operating temperature and consequently, the energy consumption and possible polymer chain degradation. In this case, the solvent expands the coil and increases the chain mobility behaving like a polymer melt where individual chains are sufficiently separated. However, at higher concentrations the solution becomes semidilute and eventually entangled depending on the degree of overlap of adjacent coils and their molar mass [33]. To sum up, the sorption of CO2 in PS/solutions is the responsible of the plasticisation which entails physico-chemical changes of their properties. The study of the properties shifts as a function of pressure, temperature and concentration will determine the morphology of the recovered Polystyrene and they are very useful for the design of the process. According to this issue, the first aim of the section is to obtain viscosity data for PS/Limonene solutions in the presence of CO 2 using a high pressure quartz crystal viscometer. These measurements concern essentially solvent-rich one-phase systems, which mean that they are focused on the influence of a known amount of CO2 on the homogeneous polymeric solution. The pressure and temperature dependence of the viscosity was analysed using the Arrhenius equation and through this theory the activation energy for viscous flow was investigated. Finally the demixing points of the solutions were determined using the experimental viscosity data. On the other hand, the second objective of this Chapter is to analyze the interfacial tension of Polystyrene, in dilute solution with carbon dioxide at high pressure. The solvent allows to work at lower temperature than the used traditionally, where temperature exceeds the polymer melting point. IFT can affect the dispersion, fluid dynamics and mass-transfer rates during processes, and its knowledge it is necessary for the design and modelling of multiphase systems with supercritical fluids. 6.2. Viscosity of mixtures CO2/Limonene/Polystyrene The viscosity of polymer solutions in the presence of CO 2 depends on multiple factors. In this work, we varied: pressure, temperature and PS concentration, while 157 Ternary Systems: Properties the solvent (Limonene) and the molecular weight distribution of the polymer were kept constant. Viscosity data were obtained for solutions of PS in Limonene at polymer concentrations of 0.025 to 0.400 g PS/ml. The viscosity measurements were carried out in the temperature range from 20 to 40ºC and at pressures between atmospheric and 50 bar. Higher pressures were not performed because there could be phase separation [34]. The concentration of CO2 in the mixtures was in the range 0–35 wt %. Figure 6.1 shows the viscosity of PS solutions in Limonene at various temperatures, pressures and polymer concentrations. 0.6 C0: 0.05 g PS/ml Lim Viscosity (mPa·s) 0.5 Patm 0.4 0.3 P: 6.89 bar 0.2 P: 13.79 bar 0.1 P: 20.68 bar 0.0 Viscosity (mPa·s) 6 C0: 0.10 g PS/ml Lim 5 P: 6.89 bar 4 3 P: 27.58 bar 2 1 0 Viscosity (mPa·s) Patm P: 34.47 bar P: 41.37 bar P: 48.26 bar C0: 0.30 g PS/ml Lim Patm 20 15 P: 6.89 bar 10 P: 13.79 bar P: 27.58 bar P: 48.26 bar 5 0 20 25 30 Temperature (ºC) 158 35 40 Chapter 6 Figure 6.1. Viscosity data showing the effect of variables: pressure, temperature and PS concentration in the initial solutions. The effect of pressure (from atmospheric pressure to 48.26 bar) is observed at constant concentration in each graph as it is shown with the different lines pointing to the arrows and labels. The effect of temperature at constant concentration is shown in the abscissa axis and the PS concentration in Limonene is noticed along the different graphs: (a) C0: 0.05 gPS/ml Limonene; (b) C0: 0.10 gPS/ml Limonene; (c) C0: 0.30 gPS/ml Limonene. The reproducibiliy of the experimental viscosity measurements was determined by comparing the results from two independent runs carried out under identical conditions. In these experiments the viscosity was similar with an average deviation of 0.508 mPa·s, indicating good reproducibility. 6.2.1. Selection of the parameters for the study of the viscosity Initially, a design of experiments (DOE) was carried out to determine if the selected variables showed a significant influence on the viscosity. These data (Figure 6.1.) were used to determine the most relevant and/or significant factors influencing viscosity. There are different ways to design a test, but the most common is a full factorial experiment. In this case, we selected the simplest factorial designs which involve the three factors at two levels and three central points [35]. Next, statistical analysis of variance (ANOVA) was applied to decide whether the effect of the input factors was significant. The range of the input factors was selected according to those used in recycling of PS wastes by dissolution and supercritical CO 2 [36]. Also the working range was selected considering the phases stability which was previously observed in a high pressure view cell [37]. The statistical analysis of the experimental results was made using commercial software package (Statgraphics Plus 5.1 Manugistics, Inc. Rockville, MD, USA). The test of statistical significance (p-value) was determined accordingly to the total error criteria giving a confidence level of 95%. Figure 6.2 illustrates the Pareto chart which shows the influence of the studied variables on the viscosity. For a significance level of 5%, the vertical line indicates the minimum statistically-significant effect. The horizontal bar lengths are proportional to the degree of significance for each effect, and every effect that exceeds the vertical line is judged significant. According to Figure 6.2, all the selected variables have a statistically significant influence over the response. Also the effects of combinations of pressure/temperature, pressure/concentration or temperature/ concentration are shown to be statistically significant over the viscosity. The filled black bars show a negative influence over the response which means that an increase in the variable 159 Ternary Systems: Properties acts to decrease the viscosity of the mixture. By contrast, the striped bars show a positive influence over the viscosity which means that variable increases the viscosity (this effect was only apparent in the case of concentration). The selected variables will be investigated in the subsequent section according to their statistically significant magnitude. C:Concentration CC A:Pressure AC B:Temperature BC BB AB AA + - 0 10 20 30 40 Standarized Effects Figure 6.2. Pareto chart of the standardized effects on Viscosity. The vertical line defines the statistically significant variables, in this case all the studied. A: pressure, B: temperature, C: concentration, AB: combined effect between pressure and temperature, AC: combined effect between pressure and concentration, BC: combined effect between temperature and concentration. 6.2.2. Concentration effect on viscosity Polystyrene is a solid at room temperature, whose melt viscosity is in the order 10 5 Pa·s at 175 ºC [38, 39] so the presence of the polymer certainly increases the viscosity of the solvent. Most theories about viscosity of polymer solutions are formulated around dilute solutions but, for real polymer solutions, the study of viscosity is often more complex than simple theory predicts. In this work, all the polymer solutions lie on the semidilute entangled and concentrated entangled regimes following Graessley’s classification [40]. The presence of a small amount of polymer increases significantly the viscosity of the Limonene solutions entailing an important asymmetry between the components responsible of thermodynamic and dynamics changes that distance from the ideal behaviour and makes very difficult its accurate prediction. Viscosity data are typically presented in dimensionless form, relating the ratio between the solution and solvent viscosities through the relative viscosity (rel). 160 [6.1] Chapter 6 where sol in [mPa·s] is the viscosity of the PS/Limonene/CO 2 solution and solvent in [mPa·s] is the viscosity of the solvent under identical working conditions. The parameter rel is a function of the pressure, temperature and initial PS concentration; it should be always above 1 because of the presence of the polymer which will increase the viscosity of the solution. The influence of pressure, temperature and concentration on the relative viscosity have been studied (Figure 6.3), but in this section only the concentration effect will be discussed. The data show an almost linear increase of rel when concentration increases, especially at low pressure. However, at higher pressure, the effect of concentration on viscosity becomes less important and it does not contribute linearly. This effect is only significant from 0.10 g/ml because the presence of CO 2 leads to a dilution effect, as well as the plasticisation of the polymer which will be explained below. The fact that the concentration effect prevails over the temperature is also shown as in the Pareto chart (Figure 6.3). The values of viscosity of the PS/Limonene solutions are lower than the values reported in the literature for methylcyclohexane and toluene [1, 41], which suggests a lower chain extension of the polymer in Limonene due to the thermodynamic qualities of the solvent [42] that will influence on the mobility of the entangled chains and consequently in the values of viscosity. (a) T: 25 ºC Relative viscosity 20 15 10 5 0 0.4 0.3 40 0.2 30 0.1 Concentration (g/ml) 20 0 10 Pressure (bar) Figure 6.3. Three-dimensional plot of the relative viscosity (rel) in a range of pressure between 10 and 50 bar, concentration from 0.05 g PS/ ml Limonene to 0.40 g PS/ ml Limonene and at three different temperatures: (a) T: 25ºC, (b) T: 30ºC and (c) T: 40ºC. 161 Ternary Systems: Properties (b) T: 30 ºC Relative Viscosity 20 15 10 5 0 0.4 0.3 40 0.2 30 0.1 20 0 Concentration (g/ml) 10 Pressure (bar) (c) T: 40 ºC Relative viscosity 20 15 10 5 0 0.4 0.3 40 0.2 30 0.1 Concentration (g/ml) 20 0 10 Pressure (bar) Figure 6.3 (Continued). Three-dimensional plot of the relative viscosity (rel) in a range of pressure between 10 and 50 bar, concentration from 0.05 g PS/ ml Limonene to 0.40 g PS/ ml Limonene and at three different temperatures: (a) T: 25ºC, (b) T: 30ºC and (c) T: 40ºC. 6.2.3. Pressure effect on viscosity The influence of pressure on the viscosity at constant temperature and PS concentration (Figure 6.4) shows that when pressure increases, the viscosity decreases following a logarithmic behaviour that becomes less pronounced at higher values of temperature and concentration. Since the viscosity decreases when pressure increases, a good correlation can be obtained using an Arrhenius-type equation [43]: [6.2] where is the viscosity in [mPa·s], A in [mPa·s], B in [mPa·s] and C in [bar -1] are three constants depending on the temperature, and p is the CO 2 pressure in [bar]. However there is not a perfect fit between the Arrhenius-type equation and the experimental data (Figure 6.4). In this case, ln does not decrease linearly when 162 Chapter 6 pressure increases and other effects should be considered to describe the non linear behaviour. When pressure increases two opposite influences occur: the plasticising effect due to the dissolution of CO2 inside the PS/Limonene solutions and the compression of the solutions as a consequence of the applied pressure. 50 (a) Temperature: 25ºC Viscosity (mPa·s) 45 2 40 C0: 0.2 g PS/ml Lim. R : 0.88 35 C0: 0.3 g PS/ml Lim. R : 0.87 30 C0: 0.4 g PS/ml Lim. R : 0.86 2 2 25 20 15 10 5 40 0 (b) Temperature: 30ºC Viscosity (mPa·s) 35 2 C0: 0.2 g PS/ml Lim. R : 0.97 30 C0: 0.3 g PS/ml Lim. R : 0.90 25 C0: 0.4 g PS/ml Lim. R : 0.89 2 2 20 15 10 5 0 (c) Temperature: 40ºC Viscosity (mPa·s) 30 2 C0: 0.2 g PS/ml Lim. R : 0.97 2 25 C0: 0.3 g PS/ml Lim. R : 0.93 20 C0: 0.4 g PS/ml Lim. R : 0.88 2 15 10 5 0 0 10 20 30 40 50 60 70 Pressure (bar) Figure 6.4. Viscosity data showing the variation of the viscosity as a function of pressure at three different temperatures (a) T: 25ºC; (b) T: 30ºC; (c) T: 40ºC and at different concentration as it is indicated by the markers: (▲) C0: 0.2 g PS/ml Limonene; () C0: 0.3 g PS/ml Limonene; () C0: 0.4 g PS/ml Limonene. Error bars represent standard deviation based on two measurements. The first effect is the increase of the CO 2 density along with the CO2 sorption into the polymer and/or the polymeric solution causing the plasticisation effect [44]. The smaller CO2 molecules placed among the entangled polymeric chains decrease their interactions and contribute to a decrease in viscosity by increasing the spatial 163 Ternary Systems: Properties distance among them (i.e. the free volume fraction of the polymer). This effect becomes more important at higher pressure and lower temperature, because the solubility of CO2 in the liquid phase increases as the pressure increases, whilst it decreases as temperature increases [45, 46]. An accurate prediction of CO2 sorption inside the solution is complex to obtain so experimental data are necessary to check the estimation. The amount of CO2 absorbed in the PS/Limonene solution causes the plasticisation and consequently the decrease of viscosity. Based on the study of other properties in previous work [37], the viscosity depression was connected with the amount of CO2 inside the solution (Figure 6.5). Pressure (bar) 0 5 10 15 20 Viscosity (mPa·s) 35 25 30 35 (a) Temperature:25ºC C0: 0.05 g/ml 30 C0: 0.10 g/ml 25 C0: 0.20 g/ml C0: 0.30 g/ml 20 C0: 0.40 g/ml 15 10 5 0 0 5 10 15 20 25 30 35 25 30 35 % wt CO2 Pressure (bar) 0 5 10 15 20 Viscosity (mPa·s) 35 (b) Temperature:30 ºC C0: 0.05 g/ml 30 C0: 0.10 g/ml 25 C0: 0.20 g/ml C0: 0.30 g/ml 20 C0: 0.40 g/ml 15 10 5 0 0 5 10 15 20 25 30 % wt CO2 Pressure (bar) 0 5 10 15 Viscosity (mPa·s) 30 20 25 30 35 (c) Temperature:40 ºC C0: 0.05 g/ml 25 C0: 0.10 g/ml 20 C0: 0.20 g/ml 15 C0: 0.40 g/ml C0: 0.30 g/ml 10 5 0 0 5 10 15 20 25 % wt CO2 Figure 6.5. Influence of CO2 sorption on the viscosity of solutions at 25ºC (a); 30ºC (b) and 40ºC (c) as a function of concentration (■) C0: 0.05 gPS/ml Limonene; (○) C0: 0.10 gPS/ml Limonene; (▲) C0: 0.20 gPS/ml Limonene; () C0: 0.30 gPS/ml Limonene; ()C0: 0.40 gPS/ml Limonene. As shown in Figure 6.5, the CO2 solubility in PS/Limonene solutions increases with increasing pressure and decreases with increasing temperature. The effect of 164 Chapter 6 temperature on CO2 sorption is observed through the length of the lines obtained as a function of temperature and PS concentration. At 25 ºC the CO2 sorption inside the solution reaches around 30% while at 40 ºC it decreases up to 20%.The effect of CO2 pressure is diminished at higher temperatures because less CO 2 is dissolved in the solution [47]. Also it should be pointed out that the reduction in the viscosity is smaller at higher CO2 concentration because the pressure effect starts to dominate at higher pressures [48]. CO2 sorption is also affected by the polymer initial concentration, when PS concentration increases CO2 solubility in solutions decreases and subsequently viscosity depression was less pronounced [49]. The second effect of pressure over viscosity is that the density of the CO2 and the bulk increases, which decreases both the molecular mobility of the polymer chains and the free volume between the chains [50]. This increases the internal frictional forces and, as a consequence, the flow resistance. Lucas suggested that changes in the pressure resulted in an increase in viscosity (when no CO 2 is added) and those modifications can be estimated from a theoretical point of view [51, 52]. The estimation method is based on the correlation of polar and nonpolar liquids, where the ratio between the viscosity of the liquid at the studied pressure and the viscosity of the saturated liquid at the vapour pressure was obtained as a function of the Pr that is defined as: [6.3] where P is the studied pressure expressed in [bar], Pv is the vapour pressure of Limonene [53] and Pc is the critical pressure of Limonene [51]. Also the ratio between viscosities is influenced by the acentric factor and the working temperature. Following the method described by Lucas [52] the effect of pressure on the solvent was calculated theoretically (Figure 6.6 solid lines). The effect of CO2 pressure on the viscosity of Limonene (Figure 6.6 dashed lines) was also predicted theoretically according to Grunberg-Nissan method [54] because it is widely recommended for liquid mixture viscosity at reduced temperatures below 0.7. The interaction parameter was obtained according to the procedure established by Isdale [55] instead of regression because this technique yields quite acceptable estimates of low-temperature liquid mixture viscosities. The effect of hydrostatic pressure and CO 2 pressure on the viscosity of Limonene clearly shows the plasticisation effect as viscosity decreases dramatically as the CO 2 pressure increases almost linearly. This behaviour has been discussed by several authors [47, 48]. Nevertheless, in the absence of CO2, the viscosity of Limonene increases slightly while the pressure increases. 165 Ternary Systems: Properties Viscosity (mPa·s) 1.0 0.8 0.6 0.4 Viscosity (mPa·s) 0.2 (a) Temperature: 25 ºC Limonene Limonene/CO2 0.0 0.8 0.6 0.4 Viscosity (mPa·s) 0.2 (b) Temperature: 30 ºC Limonene Limonene/CO2 0.0 0.8 0.6 0.4 0.2 (c) Temperature: 40 ºC Limonene Limonene/CO2 0.0 0 5 10 15 20 25 30 35 40 45 Pressure (bar) Figure 6.6. Predicted theoretically viscosity of Limonene as a function of hydrostatic pressure and temperature (solid lines) and predicted theoretically viscosity of Limonene when the pressure is provided by CO 2 (dashed lines). Clearly, the main effect of CO2 pressure on viscosity is a depression as a consequence of the plasticisation effect but also it is clear that the hydrostatic pressure is important when the working pressure is high enough, but in general, this effect is outweigh by the plasticisation caused by CO2 [46]. 6.2.4. Temperature effects on viscosity The influence of temperature on the viscosity at constant pressure (each graph) and PS concentration (different markers) is shown in Figure 6.7. As the temperature of the solution increases (top axis), the viscosity decreases. When the solutions are heated the cohesive and attraction forces between the molecules are reduced while the kinetic energy of the molecules increases which allows the free movement and consequently reduces the viscosity of the mixture. 166 Chapter 6 Temperature (ºC) 4.0 40 35 30 25 20 3.6 3.2 3.2 ln (Viscosity) ln (Viscosity) 3.6 2.8 2.4 2.0 (a) Patm C0=0.1 g/ml 1.6 C =0.2 g/ml 0 1.2 C0=0.3 g/ml 0.8 C0=0.4 g/ml 0.0031 Temperature (ºC)25 30 40 35 C0=0.2 g/ml C0=0.3 g/ml 2.8 C0=0.4 g/ml 2.4 2.0 1.6 1.2 0.0032 0.0033 0.0034 0.0031 -1 20 (c) P:13.79 bar C0=0.1 g/ml 3.2 C0=0.2 g/ml 2.8 C0=0.3 g/ml 2.4 0.0034 Temperature (ºC) 25 3.6 ln (Viscosity) ln (Viscosity) 3.6 30 0.0033 1/Temperature (K ) Temperature (ºC) 35 0.0032 -1 1/Temperature (K ) 40 20 (b) P:6.89 bar C0=0.1 g/ml C0=0.4 g/ml 2.0 1.6 40 35 30 25 20 3.2 (d) P:27.58 bar C0=0.1 g/ml 2.8 C0=0.2 g/ml 2.4 C =0.3 g/ml 0 2.0 C0=0.4 g/ml 1.6 1.2 0.8 1.2 0.4 0.8 0.0 0.0031 0.0032 0.0033 0.0034 -1 1/Temperature (K ) 0.0031 0.0032 0.0033 0.0034 -1 1/Temperature (K ) Figure 6.7. Influence of temperature on the logarithm of viscosity at different concentrations: 0.1 g PS/ml Lim (○), 0.2 g PS/ml Lim (▲), 0.3 g PS/ml Lim (), 0.4 g PS/ml Lim (). The influence of pressure is observed in each graph: (a) atmospheric pressure; (b) p: 6.89 bar; (c) p: 13.79 bar and (d) p: 27.58 bar. The viscosity dependence of the temperature can also be described using an exponential Arrhenius-type equation. [6.6] where B is a constant in [mPa·s], Ea is the flow activation energy in [J/mol] , R is the gas constant [8.3145 J/K·mol] and T is the temperature in [K]. The activation energies of the PS solutions were evaluated from the slope of ln versus 1/T plots at each pressure and concentration. The calculated activation energy ranged from 1.5 to 43 kJ/mol (Figure 6.9). 167 Ternary Systems: Properties 50 C0: 0.1 g/ml 45 C0: 0.2 g/ml 40 C0: 0.3 g/ml Ea (kJ/mol) 35 30 25 20 15 10 5 0 0 10 20 30 40 50 60 Pressure (bar) Figure 6.8. Effect of pressure on the flow activation energy (Ea). The markers represent the influence of PS concentration on the flow activation Energy: (○) 0.1 g PS/ml Lim; (▲) 0.2 g PS/ml Lim; () 0.3 g PS/ml Lim. The energy of activation is expected to be related to the latent heat of vaporization of the solvent, since the removal of a molecule from the surroundings of its neighbours forms a part of both processes. In this case, the Limonene heat of vaporization is around 40 kJ/mol [56]. Nevertheless the obtained data showed lower values than those reported in the literature. The CO 2 is less viscous than the Limonene and obviously less viscous than the Limonene/PS solutions. When the CO2 is absorbed into the solutions, they become homogeneous expanded solutions whose viscosity is decreased and therefore the Ea is lower than the heat of vaporization of Limonene. The fact that in some particular case (C0: 0.3 g/ml – P: 50 bar) Ea reaches higher values is due to the presence of PS. The Ea data obtained were in the same order of magnitude as other PS solutions using different solvents [1, 41, 57]. In the case of PS/Toluene or PS/Methylcyclohexane solutions, Ea was in the range of 7-10 kJ/mol and the authors considered the energy of activation essentially unchanged with the CO 2 addition in the investigated range. Nevertheless, in the case of PS/Limonene solutions the activation energy was higher because our measurements were conducted with samples of higher molecular weight and at higher concentration. Generally, more energy is required to enhance the polymeric chain mobility when they present higher molecular weight, indicating an increasing effective size for the flow unit [58]. Flow activation energy increases when pressure increases because the polymeric chains are restrained to move freely. The increase of CO 2 density is the responsible 168 Chapter 6 to press the solution which will make necessary a higher energy contribution to shift the polystyrene chains. Also the flow activation energy increases when concentration is higher because the polymer chains are not so accessible as in the case of dilute solutions. Higher energy should be applied to enhance the movement of the chains which are more strongly interacting. The cohesive forces in solids are at a maximum because the molecules are closer than in the case of the liquids, so the presence of the polymer increases significantly the activation energy. 6.2.5. Pressure and temperature combined effects on viscosity Although all the combined effects were significant in influencing the viscosity, only Pressure-Temperature was studied because it can be easily related through the density. If viscosity is dependent upon pressure and temperature, it should be easier to consider both effects by considering CO 2 density. Density will comprise the effect of temperature and pressure, so it will be possible to check the most relevant effect upon the viscosity as described previously by Pareto chart (Figure 6.2). The effect of density on the viscosity (Figure 6.9) shows that viscosity decreases when the density of CO2 increases. Higher values of density correspond to higher values of pressure and lower values of temperature. At higher density the solubility of CO2 into the PS/Limonene solutions increases which causes a dispersion and dilution of the large polymeric molecules. 50 C0:0.2 g/ml 45 C0:0.3 g/ml Viscosity (mPa·s) 40 C0:0.4 g/ml 35 30 25 20 15 10 5 0 0 25 50 75 100 125 150 175 200 3 Density (kg/m ) Figure 6.9. Influence of CO2 density on viscosity. Experimental data (markers) represent the viscosity at C0: 0.2 g PS/ml Lim (▲); C0: 0.3 g PS/ml Lim () ; C0: 0.4 g PS/ml Lim ().Lines show the correlation of viscosity data vs density at C0: 0.2 g PS/ml Lim (—); C0: 0.3 g PS/ml 169 Ternary Systems: Properties Lim () ; C0: 0.4 g PS/ml Lim (…). An empirical model was developed by Seifried and Temelli [59] to correlate the viscosity of fish oils triglycerides to temperature and pressure. In this work, some modifications were proposed to obtain more accurate data in the case of the PS/Limonene solutions in presence of CO 2. The following equations were used to correlate the experimental data obtained. [6.5] [6.6] [6.7] [6.8] where (P,T) is the experimental viscosity as a function of pressure and temperature in [mPa·s]. The term (P,T) explains the asymptotic behaviour of the solution at high pressure and temperature expressed in [mPa·s]. It was amended to include the effect of pressure previously not described by Seifried and Temelli [59]. The term 0 (T) represents the viscosity at atmospheric pressure in [mPa·s], for this reason only the temperature affects the term since the pressure is constant (1 atm). A4 and A5 can be obtained from the correlation of experimental data shown by Clará et al. [56]. The reported data were obtained at atmospheric pressure as a function of temperature. Nevertheless, it is important to emphasise that the presence of the PS will modify the cited values. Finally, k (P,T) shows the exponential decline of viscosity as a function of pressure and temperature as explained previously. The described model is totally empirical with the calculated parameters (Table 6.1) describing the experimental data accurately over the range of the studied conditions. Table 6.1. Model parameters for correlation of viscosity of PS/Limonene solutions in presence of CO2. C0 A1 A2 A3 A4 (g/ml) [mPa·s] [mPa·s·K-1] [mPa·s·bar-1] [mPa·s] A5 [K-1] 875.770 A6 A7 [bar-1] [bar·K-1] 0.05 -0.285 0.002 -0.115 1.680E-02 0.651 -0.090 0.10 9.204 -0.015 -0.848 9.900E-03 3500.000 0.440 -0.090 0.20 20.236 -0.049 -1.231 6.406E-05 3544.232 0.197 -0.096 0.30 61.788 -0.173 -1.638 3.174E-05 3925.024 0.120 -0.024 0.40 54.295 -0.092 -15.878 2.470E-02 2096.700 0.058 -0.005 Figure 6.10 shows the experimental (markers) and correlated (lines) viscosity data as a function of concentration (each graph), pressure and temperature. In all cases there was a good fit between the experimental and the correlated data which allows to interpolate new data in the experimental range studied and to extrapolate to 170 Chapter 6 moderately higher pressures and temperatures. The increase of the concentration of PS would not imply a worsening of the data because the parameters were obtained from the fitting of five different concentrations covering a wide range of working conditions. The modification of the semi-empirical equations allows the improvement of the correlation overcoming the non ideality that could provide the presence of the polymer. 5.0 10 (a) C0: 0.1 gPS/ml Lim T: 25 ºC T: 30 ºC T: 40 ºC Viscosity (mPa·s) 4.0 3.5 8 3.0 2.5 2.0 1.5 1.0 0.5 7 6 5 4 3 2 1 0.0 0 0 10 20 30 40 50 60 70 80 0 Pressure (bar) 20 14 20 30 40 10 8 6 4 2 60 70 80 (d) C0: 0.4 gPS/ml Lim T: 25 ºC T: 30 ºC T: 40 ºC 24 12 50 Pressure (bar) 27 Viscosity (mPa·s) 16 10 30 (c) C0: 0.3 gPS/ml Lim T: 25 ºC T: 30 ºC T: 40 ºC 18 Viscosity (mPa·s) (b) C0: 0.2 gPS/ml Lim T: 25 ºC T: 30 ºC T: 40 ºC 9 Viscosity (mPa·s) 4.5 21 18 15 12 9 6 3 0 0 0 5 10 15 20 25 30 35 40 Pressure (bar) 45 50 55 60 0 5 10 15 20 25 30 35 40 45 50 55 60 Pressure (bar) Figure 6.10. Viscosity of PS/Limonene solutions as a function of pressure, temperature at 25 ºC (■), 30 ºC (●) and 40 ºC (▲) and concentration (a) C0: 0.10 g PS/ml Lim; (b) C0: 0.20 g PS/ml Lim; (c) C0: 0.30 g PS/ml Lim; (d) C0: 0.40 g PS/ml Lim. The correlated data calculated according to eq. (6.5) are shown at 25 ºC (—), 30 ºC () and 40 ºC (…). Clearly, the pressure and temperature or density effect on viscosity is an indirect measurement of the CO2 sorption which has been shown as the real cause of changes in viscosity of the polymeric solutions. CO2 is a small molecule that can easily penetrate between the polymer chains causing a dilution effect and increasing their free volume [39, 46]. As gas content increases, viscosity decreases because the CO2 sorption inside the solution increases while pressure increases and temperature decreases. 171 Ternary Systems: Properties 6.2.6. Demixing determination from viscosity measurements Pressure-induced phase separation is a relatively new technique that provides an alternative path to enter the unstable regions in polymer solutions [22]. The viscosity can be used to predict the pressure of vanishing, it means, the free energy density of CO2 becomes closer to that of the polymer-solvent phase and the viscosity decreases. When gas pressure increases the concentration of CO 2 in the polymersolvent phase increases promoting a decrease of viscosity, since the two phases in contact become more similar. In the late 1930s and 1940s Mark, Houwink and Sakurada (MHS) arrived at an empirical relationship between the molecular weight and the intrinsic viscosity [6.9] Where k is a constant in [cm3·g-1], MW is the molecular weight of the polymer and a is called Mark-Houwink-Sakurada exponent and it provides a measure of the chain rigidity. According to the experimental viscosity data obtained and knowing the concentration of the solutions as well as the molecular weight of the polymer, the MHS parameter can be calculated from the previous equation. The value of a is thought to be indicative of the solvent power. As a polymer chain is a good solvent of itself, the polymer chains are much more stretched and sometimes the chain form molecular entanglements as in the case of this work. As explained, the viscosity of PS/Limonene solution decreases in the presence of CO 2 as a consequence of the plasticisation effect. In previous work it was noticed that polymer-solvent phase separation could be accomplished if CO 2 is added as antisolvent [34]. In compressible fluids, without changing the fluid itself, solvent quality is altered by increasing the pressure inducing phase separation in polymer solutions [60]. The parameter a could reflect the change in the solvent power of Limonene when CO2 is added. Generally, a is 0.5 in the case of theta solvents, while in good solvents a is 0.8 [61]. When PS is dissolved in Limonene at atmospheric pressure and 25ºC, a is around 0.35 which is indicative of the poorer solvent power of the terpene. The parameter a was calculated in the working range of pressure, temperature and concentration (Figure 6.11). 172 Chapter 6 0.36 0.34 0.32 0.30 a 0.28 0.26 T: 25 ºC 0.24 T: 30 ºC T: 40 ºC 0.22 0.20 0 5 10 15 20 25 Pressure (bar) Figure 6.11. Influence of pressure and temperature on parameter a. Markers represent the different temperatures: (■) 25 ºC, (○) 30 ºC and (▲) 40 ºC. The viscosity exponent decreases with increasing pressure (Figure 6.11), that is with increasing the solubility of CO2 in the solution, revealing that the solvent goodness lowers as the CO2 concentration in the mixture increases. When Limonene is fully miscible in CO2, its solvent power with regard to the polymer will become poorer and the parameter a will be close to 0. On the basis of an extrapolation of a, it should be possible to determine the critical pressure at which phase separation will appear (Table 6.2). Table 6.2. Critical pressure of the PS/Limonene solutions as a function of temperature based on the determination of viscosity. Temperature (ºC) Pressure (bar) 24 63.29 29 71.29 34 78.39 The viscosity depends on the CO2 pressure to a greater degree in higher concentration solutions, presumably because CO 2 is an antisolvent for PS that reduces the favourability of interactions between the polymer and solvent. It was proposed that the solvent power of Limonene/CO 2 mixtures decreased with increasing CO2 pressure, and this increased the importance of polymer–polymer interactions and thereby reduced the entanglement density [47]. Finally, the influence of pressure, temperature and concentration on the viscosity of PS/p-Cymene solutions was also studied. Results have been depicted in Figure 6.12. 173 Ternary Systems: Properties 6 (a) C0: 0.3 g PS/ml p-Cymene P: 6.89 bar P: 20.68 bar P: 27.58 bar P: 34.47 bar P: 48.26 bar Viscosity (mPa·s) 5 4 3 2 1 0 (b) C0: 0.4 g PS/ml p-Cymene Patm P: 6.89 bar P: 13.79 bar P: 20.68 bar P: 27.58 bar Viscosity (mPa·s) 40 30 20 10 0 20 25 30 35 40 Temperature (ºC) Figure 6.12. Viscosity data showing the effect of variables: pressure, temperature and concentration of PS in the initial solutions of PS/p-Cymene at (a) C0: 0.3 g PS/ml p-Cymene and (b) C0: 0.4 g PS/ml p-Cymene. According to Figure 6.12, the viscosity decrease due to the increase of temperature and pressure is confirmed, as well as the viscosity increase as a consequence of the polymer concentration increase. In conclusion, the viscosity of PS/Terpene solutions was studied with the aid of a quartz viscometer in the presence of high pressure CO 2. All the studied variables: pressure, temperature and concentration presented statistical significant over the viscosity. It was checked that the viscosity of the solutions decreases on increasing pressure, independently of the concentration and temperature. Despite the reduction of the viscosity because of the plasticisation effect of CO2 the increase of density of the bulk due to the application of pressure to an incompressible fluid should be also considered. When temperature increases, the viscosity of PS/Terpenes solutions decreases and the flow activation energy (Ea) could be calculated. The values of Ea showed that higher energy is required when the 174 Chapter 6 pressure of CO2 and the concentration of PS increase. The changes of viscosity could be used to determine the solvent-solute interaction in order to determine the stability of each phase by means of the Mark-Houwink-Sakurada equation. The results determined the critical pressure which the phase separation between the polymer and the solvent by CO2 (antisolvent) could be appeared. 6.3. Interfacial tension of CO2/Limonene/PS and CO2/p-Cymene/PS The effect of pressure, temperature, concentration and type of solvent on the interfacial tension (IFT) of PS solutions in the presence of CO 2 was studied. IFT data were obtained for solutions of 0.05 and 0.20 g PS/ ml terpene. The IFT measurements were carried out in the temperature range from 30 to 40 ºC and at pressures from atmospheric up to the vanishing point. The accuracy of the experimental IFT measurements was determined by comparing the results from four independent runs carried out under identical conditions. In these experiments the IFT were similar with a deviation of 0.0373 mN/m, indicating that reproducibility of the data was also good. The influence of pressure, temperature and the initial concentration the solvent over the IFT of CO 2/pCymene/PS and CO2/Limonene/PS systems is presented in Figure 6.13. 35 35 (a) C0=0.05 g PS/ml p-Cymene 25 20 15 30 ºC 40ºC 25 20 15 10 10 5 5 0 0 0 10 20 30 40 50 60 70 80 90 0 Pressure (bar) 35 20 (mN/m) 20 15 30 40 50 60 70 80 90 Pressure (bar) (d) C0=0.20 g PS/ml Limonene 30 30 ºC 40ºC 25 10 35 (c) C0=0.20 g PS/ml p-Cymene 30 (mN/m) (b) C0=0.05 g PS/ml Limonene 30 30 ºC 40ºC (mN/m) (mN/m) 30 30 ºC 40ºC 25 20 15 10 10 5 5 0 0 0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90 Pressure (bar) Pressure (bar) Figure 6.13. Influence of pressure, temperature and initial concentration on IFT of PS dissolved in Cymene (a, c) and Limonene (b, d). 175 Ternary Systems: Properties In these Figures it can be seen that the interfacial tension decreases linearly as the pressure of CO2 increases. This decrease is generally attributed to two phenomena [62-64]: as pressure increases, the free energy density of CO 2 becomes closer to that of the polymer-solvent phase and the interfacial tension decreases; and as gas pressure increases, the concentration of CO 2 in the polymer-solvent phase increases thus further promoting a decrease of interfacial tension since the two phases in contact become more similar. On the other hand, in both systems the pressure and the interfacial tension value at which the interfacial tension vanishes (both reported in Table 6.3) slightly increase with the temperature, especially at higher concentration. Table 6.3. Effect of solvent, temperature and initial concentration of PS dissolution on the critical pressure obtained from experimental IFT extrapolation. Solvent C0 (g PS /ml) Temperature (ºC) Slope (mN/m·bar) Intercept (mN/m) Critical Pressure (bar) Solvent p- Cymene 0.05 30 0.3766 28.7824 76.42 0.20 40 0.3320 27.9334 84.31 30 0.3756 27.0665 72.06 40 0.3337 27.7747 83.24 Limonene C0 (g PS/ml) 0.05 Temperature (ºC) 30 40 Slope (mN/m·bar) 0.3233 0.3299 Intercept (mN/m) 26.1692 28.1363 Critical Pressure (bar) 80.94 85.29 0.20 30 40 0.4268 0.3289 30.9898 27.4318 72.61 83.40 The effect of temperature on the interfacial tension requires a more complex explanation, since the temperature produces two opposite effects: i) as temperature increases the average kinetic energy of the molecules also do. So, an increase of the temperature produces a decrease on the interfacial tension, since, the interfacial tension accounts the energy required to create a new surface at the interface between two phases [65]; ii) as the temperature increases the CO2 density decreases, and therefore the solubility in CO2 of the polymer/cosolvent system decreases. In Figure 6.12 and Table 6.3 it can be seen that the second effect is dominant in the experimental range since when temperature increases, it is observed an increase of both: the interfacial tension and the pressure of vanishing. Nevertheless, in the case of 0.05 gPS/ml Limonene, where IFT slope is kept constant, both effect (increase of kinetic energy and CO2 density) are similar and the effect should be study in a wider range to explain the exception. Quantitatively IFT decrease is confirmed by the negative slope of the straight lines in Figure 6.13 (values are reported in Table 6.3). While IFT is not noticeably 176 Chapter 6 affected by the PS initial concentration temperature causes a significant slope reduction. The average correlation obtained was 0.9957. These results are in good agreement with those reported by Park et al. [66] which studied the surface tension of molten PS in supercritical CO 2 at high temperature (170-210 ºC). They found that surface tension varies linearly with temperature and pressure, concluding that the effect of temperature on it becomes lower with increasing pressure. This fact implies that a temperature increase is only effective if mild pressure is used during PS processing. The possibility of correlate the interfacial tension of polymer is important for the design of further processes. There are several techniques to estimate the surface tension of pure liquids, but the most useful are empirical. Macleod-Sugden [67] proposed a correlation to calculate the surface tension of liquid mixture. That is not a simple equation relating the surface tensions of the pure components of the mixture because the composition near the surface is not the same than that of the bulk. σ 1/4= [P] (ρL- ρV) [6.10] where P is the Parachor number, a parameter temperature independent that might be estimated from the molecular structure and LV are the liquid and vapour densities, respectively. Following the previous equations the IFT of solvents was calculated, without considering the polymer concentration at 30 and 40ºC at atmospheric pressure. The results are shown in Table 6.4. Table 6.4. Predicted values and Average Absolute Deviation (AAD) of p-Cymene and Limonene IFT according to semiempirical methods. p-Cymene Harrison et al., 1998 [67] Brock and Bird, 1955[68] Miller, 1963[69] Pitzer, 1995 [70] Sastri and Rao, 1995 [71] AAD 30ºC 26.019 27.113 27.050 28.208 27.485 0.537 40ºC 25.158 26.166 26.105 27.223 26.525 0.511 Limonene Harrison et al., 1998 [67] Brock and Bird, 1955[68] Miller, 1963[69] Pitzer, 1995 [70] Sastri and Rao, 1995 [71] AAD 30ºC 25.225 26.760 26.697 26.871 27.539 0.558 40ºC 24.523 25.826 25.765 25.940 26.585 0.482 According to Table 6.4, the calculated IFT values present an average absolute deviation of 0.522 mN/m (maximum), sufficiently small to consider that the referenced methods are able to predict the interfacial tension with accuracy enough 177 Ternary Systems: Properties for most practical purposes. Also, IFT predicted values at 0 bar are in good agreement with the cited by Simões et al. [72] who presented the IFT measurements of lemon oil, whose main constituents are monoterpenes and largely Limonene (≈ 67%) [73]. Among the described equations to predict the IFT, only the developed by MacleodSugden includes the effect of pressure by means of the density of the vapour and liquid phase. This equation was used to predict Limonene/CO2 and p-Cymene/CO2 surface tension to calculate the global parameter (IFT). Park et al. [66], on the basis of their statistical investigations, proposed an equation to estimate the surface tension of the molten Polystyrene in CO2. Based on Park treatment of IFT, PS in solution can be considered as a molten polymer because of the high mobility of its chains. With this hypothesis the effect of the polymer solubilised on the IFT of the solution can be estimated. Based on the calculated IFT for the solvent/CO2 and for the PS/CO2 and considering the initial concentration of the solutions, IFT of solution value was calculated as a function of pressure, temperature and concentration, following next equation [51]. σmr=Σinxi σir [6.11] where mr is the surface tension of mixture, xi is the mole fraction of component i in the bulk phase andir is the surface tension of pure component i. The exponent r is recommended to be 1 for most hydrocarbon mixtures which would predict linear behaviour in surface tension vs. composition. Initially the mixing rule applied was the sum of the corresponding individual IFT multiplied by their mass fraction (dash line) and secondly a quadratic mixing rule was applied (continuous line), making use of an adjustable parameter depending on the initial concentration of the solution. In mixtures containing CO2 an interaction parameter is required to fit experimental data [51]. In this work, mass fraction of the solvent and the polymer was considered instead of mole fraction because PS mole fraction is difficult to calculate attending to the molecular weight distribution. The results are showed in Figure 6.14 and 6.15. As it observed in Figures 6.14 and 6.15, empirical methods to estimate IFT are more accurate at 0.05 g/ml (initial concentration) and 40º C because the error usually increases as the concentration of the component with the largest pure component surface tension increases (Polystyrene in this case). Using quadratic mixing rule at 40ºC, the IFT prediction improves slightly as it is particularly observed in Limonene at 0.20 g PS/ml. IFT prediction is useful to establish initial values or ranges, but accuracy has been demonstrated to be only acceptable in the case of low polymer concentrations because the error increases when higher concentrations are used. IFT experimental data cannot be replaced by predictions since the global aim of the research line is the development of a recycling process of PS foams by dissolution with terpenic 178 Chapter 6 solvents and during the process high polymer concentrations are preferable to work. Consequently, IFT must be checked experimentally in the case of large PS concentrations. 35 35 (a) C0:0.05 g PS/ml p-Cymene 30 T: 30ºC T: 30ºC 25 (mN/m) (mN/m) 25 (b) C0:0.20 g PS/ml p-Cymene 30 20 15 10 5 20 15 10 5 0 0 0 10 20 30 40 50 60 70 80 90 0 10 20 Pressure (bar) 40 50 60 70 80 90 Pressure (bar) 35 35 (c) C0:0.05 g PS/ml p-Cymene 30 T: 40ºC 25 (d) C0:0.20 g PS/ml p-Cymene 30 (mN/m) (mN/m) 30 20 15 10 5 T: 40ºC 25 20 15 10 5 0 0 0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90 Pressure (bar) Pressure (bar) Figure 6.14. IFT of PS/p-Cymene solutions versus the CO2 pressure. Experimental points and correlated curve. () Dash line: linear mixing rule; (—) Solid line: quadratic mixing rule. 35 35 (a) C0:0.05 g PS/ml Limonene 30 T: 30 ºC T: 30 ºC 25 (mN/m) (mN/m) 25 (b) C0:0.20 g PS/ml Limonene 30 20 15 10 5 20 15 10 5 0 0 0 10 20 30 40 50 60 70 80 90 0 10 20 Pressure (bar) 40 50 60 70 80 90 Pressure (bar) 35 35 (c) C0:0.05 g PS/ml Limonene 30 (d) C0:0.20 g PS/ml Limonene 30 T: 40 ºC T: 40 ºC 25 (mN/m) 25 (mN/m) 30 20 15 10 5 20 15 10 5 0 0 0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90 Pressure (bar) Pressure (bar) Figure 6.15. IFT of PS/Limonene solutions versus the CO2 pressure. Experimental points and correlated curve. () Dash line: linear mixing rule; (—) Solid line: quadratic mixing rule. 179 Ternary Systems: Properties To conclude, Interfacial tension of Polystyrene solutions decreases linearly with pressure and temperature while the polymer concentration has not a significant influence in the studied range. The behaviour of the PS in solution in the presence of CO2 is similar to molten polymers and by this reason, it is possible to predict easily the interfacial tension of the mixture. 6.4. Glass transition temperature of CO2/Limonene/PS and CO2/pCymene/PS A final set of experiments was carried out in order to study the influence of a cosolvent on the glass transition temperature of the polymer-solvent mixture. The influence of CO2 on the Tg of PS/Limonene solutions was assayed (Figure 6.16). According to Figure 6.16, the presence of the terpene does not produce a significant decrease regarding the Tg of the Polystyrene/CO2 binary system (Chapter 4) independently of the initial concentration of the polymer in the solution. During plasticization, polymers are swelled with a consequent increase in the free volume of the polymer, small molecules, as CO2, soak easily in through the matrix than the bigger ones. The main influence of CO2 over the PS glass transition temperature makes the Limonene plays only a secondary role on the system. By this reason, Tg reaches very close values to the obtained without solvent. 110 PS/CO2 100 Solutions C0=0.9 g PS/ml Limonene 90 Solutions C0=0.2 g PS/ml Limonene 80 Tg (ºC) 70 60 50 40 30 20 10 0 0 20 40 60 80 100 120 Pressure (bar) 140 160 180 200 Figure 6.16. Influence of CO2 pressure on the glass transition temperature of PS in Limonene solutions.(*) This work; () Initial concentration: 0.90 g PS/ml Limonene; (▲) Initial concentration: 0.20 g PS/ml Limonene. 180 Chapter 6 These results agree with the previously obtained by Stafford et al. [74] that concluded that the presence of low molecular weight substances in a polymer/CO 2 system does not affect greatly the value of the Tg observed. Also, the influence of CO2 on the Tg of PS/p-Cymene solutions was studied. Results and errors are shown in Table 6.5. Table 6.5. Influence of the pressure of CO 2 on the glass transition temperature of PS /pCymene solutions at 0.2 g PS/ ml p-Cymene. P (bar) Tg (ºC) 0 101.4±0.99 20 76.8±1.42 40.4 51.2±0.74 60.4 39.5±2.40 79.6 33.6±1.15 99.6 33.1±0.06 According to Table 6.5 is observed that Tg decreases when pressure of CO2 increases from 0 to 100 bar. From the comparison of results shown in Tables 6.5 and 4.23 (Tg of the binary system PS/CO2) is observed that the presence of p-Cymene does not decrease dramatically the Tg of the solutions. The plastising effect of CO 2 is the main responsible of the Tg drop since it is a very small molecule which penetrates better inside the polymeric chains. Results agree pretty well with the previously shown in Figure 6.16, it confirms that terpene solvents are not the main cause of the glass transition temperature decrease when there are CO 2 in the system. In conclusion, the sorption of CO2 in the Polystyrene/terpene oils solutions decreases the viscosity, interfacial tension and glass transition temperature because of the plasticisation effect of CO2. Temperature and concentration also affects the mentioned physico-chemical properties of the polymeric solutions. The vanishing of viscosity and interfacial tension can be used to determine the solvent-solute interaction in order to determine the stability of each phase and the demixing points. 181 Ternary Systems: Properties References [1] S.D. Yeo, E. Kiran, High-pressure density and viscosity of polystyrene solutions in methylcyclohexane, Journal of Supercritical Fluids, 15 (1999) 261-272. [2] H.M. Woods, M.M.C.G. Silva, C. Nouvel, K.M. Shakesheff, S.M. Howdle, Materials processing in supercritical carbon dioxide: Surfactants, polymers and biomaterials, Journal of Materials Chemistry, 14 (2004) 1663-1678. [3] O.R. Davies, A.L. Lewis, M.J. Whitaker, H. Tai, K.M. Shakesheff, S.M. Howdle, Applications of supercritical CO2 in the fabrication of polymer systems for drug delivery and tissue engineering, Advanced Drug Delivery Reviews, 60 (2008) 373387. [4] A.I. Cooper, Polymer synthesis and processing using supercritical carbon dioxide, Journal of Materials Chemistry, 10 (2000) 207-234. [5] S. Villarroya, K.J. Thurecht, A. Heise, S.M. Howdle, Supercritical CO2: An effective medium for the chemo-enzymatic synthesis of block copolymers?, Chemical Communications, (2007) 3805-3813. [6] J. Jennings, M. Beija, A.P. Richez, S.D. Cooper, P.E. Mignot, K.J. Thurecht, K.S. Jack, S.M. Howdle, One-pot synthesis of block copolymers in supercritical carbon dioxide: A simple versatile route to nanostructured microparticles, Journal of the American Chemical Society, 134 (2012) 4772-4781. [7] L. An, B.A. Wolf, Combined effects of pressure and shear on the phase separation of polymer solutions, Macromolecules, 31 (1998) 4621-4625. [8] B. Bungert, G. Sadowski, W. Arlt, Supercritical antisolvent fractionation: Measurements in the systems monodisperse and bidisperse polystyrene-cyclohexanecarbon dioxide, Fluid Phase Equilibria, 139 (1997) 349-359. [9] I.H. Lin, P.F. Liang, C.S. Tan, Precipitation of submicrometer-sized poly(methyl methacrylate) particles with a compressed fluid antisolvent, Journal of Applied Polymer Science, 117 (2010) 1197-1207. [10] I. Asghari, F. Esmaeilzadeh, Investigation of key influence parameters for synthesis of submicron carboxymethylcellulose particles via rapid expansion of supercritical CO2 solution by Taguchi method, Journal of Supercritical Fluids, 69 (2012) 34-44. [11] E. Reverchon, S. Cardea, Production of controlled polymeric foams by supercritical CO2, Journal of Supercritical Fluids, 40 (2007) 144-152. [12] X. Han, K.W. Koelling, D.L. Tomasko, L.J. Lee, Continuous microcellular polystyrene foam extrusion with supercritical CO 2, Polymer Engineering and Science, 42 (2002) 2094-2106. [13] S. Cotugno, E. Di Maio, G. Mensitieri, S. Iannace, G.W. Roberts, R.G. Carbonell, H.B. Hopfenberg, Characterization of microcellular biodegradable polymeric foams produced from supercritical carbon dioxide solutions, Industrial and Engineering Chemistry Research, 44 (2005) 1795-1803. [14] Y.C. Bae, Gas plasticization effect of carbon dioxide for polymeric liquid, Polymer, 37 (1996) 3011-3017. [15] L.J. Gerhardt, A. Garg, C.W. Manke, E. Gulari, Concentration-dependent viscoelastic scaling models for polydimethysiloxane melts with dissolved carbon dioxide, Journal of Polymer Science, Part B: Polymer Physics, 36 (1998) 1911-1918. [16] M. Onishi, Y. Nakamura, T. Norisuye, Intrinsic viscosity of polystyrene in toluene-supercritical carbon dioxide mixtures, Polymer Journal, 41 (2009) 477-481. [17] R. Liao, W. Yu, C. Zhou, Rheological control in foaming polymeric materials: I. Amorphous polymers, Polymer, 51 (2010) 568-580. 182 Chapter 6 [18] Y.W. Kho, D.S. Kalika, B.L. Knutson, Precipitation of Nylon 6 membranes using compressed carbon dioxide, Polymer, 42 (2001) 6119-6127. [19] G.H. Chong, R. Yunus, N. Abdullah, T.S.Y. Choong, S. Spotar, Coating and encapsulation of nanoparticles using supercritical antisolvent, American Journal of Applied Sciences, 6 (2009) 1352-1358. [20] M. Karimi, M. Heuchel, T. Weigel, M. Schossig, D. Hofmann, A. Lendlein, Formation and size distribution of pores in poly(ε-caprolactone) foams prepared by pressure quenching using supercritical CO 2, Journal of Supercritical Fluids, 61 (2012) 175-190. [21] L.N. Nikitin, M.O. Gallyamov, R.A. Vinokur, A.Y. Nikolaec, E.E. Said-Galiyev, A.R. Khokhlov, H.T. Jespersen, K. Schaumburg, Swelling and impregnation of polystyrene using supercritical carbon dioxide, Journal of Supercritical Fluids, 26 (2003) 263-273. [22] C. Dindar, E. Kiran, High-pressure viscosity and density of polymer solutions at the critical polymer concentration in near-critical and supercritical fluids, Industrial and Engineering Chemistry Research, 41 (2002) 6354-6362. [23] K. Wasai, G. Kaptay, K. Mukai, N. Shinozaki, Modified classical homogeneous nucleation theory and a new minimum in free energy change. 1. A new minimum and Kelvin equation, Fluid Phase Equilibria, 254 (2007) 67-74. [24] K. Wasai, G. Kaptay, K. Mukai, N. Shinozaki, Modified classical homogeneous nucleation theory and a new minimum in free energy change. 2. Behavior of free energy change with a minimum calculated for various systems, Fluid Phase Equilibria, 255 (2007) 55-61. [25] K. Nishioka, I. Kusaka, Thermodynamic formulas of liquid phase nucleation from vapor in multicomponent systems, The Journal of Chemical Physics, 96 (1992) 5370-5376. [26] J.W. Cahn, J.E. Hilliard, Free energy of a nonuniform system. III. Nucleation in a two-component incompressible fluid, The Journal of Chemical Physics, 31 (1959) 688-699. [27] S.K. Goel, E.J. Beckman, Generation of microcellular polymeric foams using supercritical carbon dioxide. I: effect of pressure and temperature on nucleation, Polymer Engineering and Science, 34 (1994) 1137-1147. [28] S.G. Kazarian, Polymer processing with supercritical fluids, Polymer Science Series C, 42 (2000) 78-101. [29] M. Sauceau, C. Nikitine, E. Rodier, J. Fages, Effect of supercritical carbon dioxide on polystyrene extrusion, Journal of Supercritical Fluids, 43 (2007) 367-373. [30] D. Liu, H. Li, M.S. Noon, D.L. Tomasko, CO2-induced PMMA swelling and multiple thermodynamic property analysis using Sanchez-Lacombe EOS, Macromolecules, 38 (2005) 4416-4424. [31] B.A. Miller-Chou, J.L. Koenig, A review of polymer dissolution, Progress in Polymer Science (Oxford), 28 (2003) 1223-1270. [32] H. Park, C.B. Park, C. Tzoganakis, K.H. Tan, P. Chen, Surface tension measurement of polystyrene melts in supercritical carbon dioxide, Industrial and Engineering Chemistry Research, 45 (2006) 1650-1658. [33] I. Teraoka, Polymer Solutions: An Introduction to Physical Properties, Wiley, New York, 2002. [34] C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. De Lucas, M.T. García, Development of a strategy for the foaming of polystyrene dissolutions in scCO2, Journal of Supercritical Fluids, 76 (2013) 126-134. 183 Ternary Systems: Properties [35] P. A. Gauglitz, F. Friedmann, S. I. Kam, W. R. Rossen, Foam generation in homogeneous porous media, Chemical Engineering Science, 57 (2002) 4037-4052. [36] C. Gutiérrez, M.T. García, I. Gracia, A. de Lucas, J.F. Rodríguez, Recycling of extruded polystyrene wastes by dissolution and supercritical CO 2 technology, Journal of Material Cycles and Waste Management, 14 (2012) 308-316. [37] C. Kwag, C.W. Manke, E. Gulari, Effects of dissolved gas on viscoelastic scaling and glass transition temperature of polystyrene melts, Industrial and Engineering Chemistry Research, 40 (2001) 3048-3052. [38] C. Kwag, C.W. Manke, E. Gulari, Rheology of molten polystyrene with dissolved supercritical and near-critical gases, Journal of Polymer Science, Part B: Polymer Physics, 37 (1999) 2771-2781. [39] J.R. Royer, Y.J. Gay, J.M. Desimone, S.A. Khan, High-pressure rheology of polystyrene melts plasticized with CO 2: Experimental measurement and predictive scaling relationships, Journal of Polymer Science, Part B: Polymer Physics, 38 (2000) 3168-3180. [40] W.W. Graessley, Polymer chain dimensions and the dependence of viscoelastic properties on concentration, molecular weight and solvent power, Polymer, 21 (1980) 258-262. [41] S.D. Yeo, E. Kiran, High-pressure viscosity of polystyrene solutions in toluene+carbon dioxide binary mixtures, Journal of Applied Polymer Science, 75 (2000) 306-315. [42] B.A. Wolf, R. Jend, Pressure and temperature dependence of the viscosity of polymer solutions in the region of phase separation, Macromolecules, 12 (1979) 732737. [43] D. Gourgouillon, H.M.N.T. Avelino, J.M.N.A. Fareleira, M. Nunes da Ponte, Simultaneous viscosity and density measurement of supercritical CO 2-saturated PEG 400, Journal of Supercritical Fluids, 13 (1998) 177-185. [44] Y. Zhang, K.K. Gangwani, R.M. Lemert, Sorption and swelling of block copolymers in the presence of supercritical fluid carbon dioxide, Journal of Supercritical Fluids, 11 (1997) 115-134. [45] S.A.E. Boyer, J.P.E. Grolier, Modification of the glass transitions of polymers by high-pressure gas solubility, Pure and Applied Chemistry, 77 (2005) 593-603. [46] M. Lee, C.B. Park, C. Tzoganakis, Measurements and modeling of PS/supercritical CO2 solution viscosities, Polymer Engineering and Science, 39 (1999) 99-109. [47] R.E. Whittier, D. Xu, J.H. Van Zanten, D.J. Kiserow, G.W. Roberts, Viscosity of polystyrene solutions expanded with carbon dioxide, Journal of Applied Polymer Science, 99 (2006) 540-549. [48] S.P. Nalawade, V.H.J. Nieborg, F. Picchioni, L.P.B.M. Janssen, Prediction of the viscosity reduction due to dissolved CO 2 of and an elementary approach in the supercritical CO2 assisted continuous particle production of a polyester resin, Powder Technology, 170 (2006) 143-152. [49] M. Lee, C. Tzoganakis, C.B. Park, Effects of supercritical CO2 on the viscosity and morphology of polymer blends, Advances in Polymer Technology, 19 (2000) 300311. [50] W. Strauss, N.A. D'Souza, Supercritical CO2 processed polystyrene nanocomposite foams, Journal of Cellular Plastics, 40 (2004) 229-241. [51] B.E. Poling, J.M. Prausnitz, J.P. O'Connell, The properties of gases and liquids, McGraw-Hill, New York, 2001. 184 Chapter 6 [52] K. Lucas, Die Druckabhängigkeit der Viskosität von Flüssigkeiten – eine einfache Abschätzung, Chemie Ingenieur Technik, 53 (1981) 959-960. [53] Y.P. Sun, Supercritical fluid technology in materials science and engineering: syntheses, properties, and applications, Marcel Dekker, 2002. [54] H.J. Kim, H.C. Moon, H. Kim, K. Kim, J.K. Kim, J. Cho, Pressure effect of various inert gases on the phase behavior of polystyrene- block -poly(n-pentyl methacrylate) copolymer, Macromolecules, 46 (2013) 493-499. [55] J.D. Isdale, G. Cartwright, J.C. MacGillivray, Prediction of Viscosity of Organic Liquid Mixtures by a Group Contribution Method, 1981. [56] R.A. Clará, A.C.G. Marigliano, H.N. Sólimo, Density, viscosity, and refractive index in the range (283.15 to 353.15) K and vapor pressure of α-pinene, d-limonene, (±)-linalool, and citral over the pressure range 1.0 kPa atmospheric pressure, Journal of Chemical and Engineering Data, 54 (2009) 1087-1090. [57] K. Liu, F. Schuch, E. Kiran, High-pressure viscosity and density of poly(methyl methacrylate) + acetone and poly(methyl methacrylate) + acetone + CO 2 systems, Journal of Supercritical Fluids, 39 (2006) 89-101. [58] E. Kiran, Z. Gokmenoglu, High-pressure viscosity and density of polyethylene solutions in n-pentane, Journal of Applied Polymer Science, 58 (1995) 2307-2324. [59] B. Seifried, F. Temelli, Viscosity and rheological behaviour of carbon dioxideexpanded fish oil triglycerides: Measurement and modeling, Journal of Supercritical Fluids, 59 (2011) 27-35. [60] E. Kiran, Polymer miscibility, phase separation, morphological modifications and polymorphic transformations in dense fluids, Journal of Supercritical Fluids, 47 (2009) 466-483. [61] D.W. van Krevelen, K. te Nijenhuis, Properties of Polymers: Their Correlation with Chemical Structure; their Numerical Estimation and Prediction from Additive Group Contributions, Elsevier Science, 2009. [62] H. Li, L.J. Lee, D.L. Tomasko, Effect of Carbon Dioxide on the Interfacial Tension of Polymer Melts, Industrial and Engineering Chemistry Research, 43 (2004) 509-514. [63] K.L. Harrison, S.R.P. Da Rocha, M.Z. Yates, K.P. Johnston, D. Canelas, J.M. DeSimone, Interfacial activity of polymeric surfactants at the polystyrene-carbon dioxide interface, Langmuir, 14 (1998) 6855-6863. [64] M.G. Pastore Carbone, E. Di Maio, G. Scherillo, G. Mensitieri, S. Iannace, Solubility, mutual diffusivity, specific volume and interfacial tension of molten PCL/CO2 solutions by a fully experimental procedure: Effect of pressure and temperature, Journal of Supercritical Fluids, 67 (2012) 131-138. [65] X. Liao, Y.G. Li, C.B. Park, P. Chen, Interfacial tension of linear and branched PP in supercritical carbon dioxide, Journal of Supercritical Fluids, 55 (2010) 386394. [66] H. Park, R.B. Thompson, N. Lanson, C. Tzoganakis, C.B. Park, P. Chen, Effect of temperature and pressure on surface tension of polystyrene in supercritical carbon dioxide, Journal of Physical Chemistry B, 111 (2007) 3859-3868. [67] D.B. Macleod, On a relation between surface tension and density, Transactions of the Faraday Society, 19 (1923) 38-41. [68] J.R. Brock, R.B. Bird, Surface tension and the principle of corresponding states, AIChE Journal, 1 (1955) 174-177. [69] D.G. Miller, Graphical methods for determining a nonlinear constant in vapor pressure equations, Industrial and Engineering Chemistry Fundamentals, 2 (1963) 68-73. 185 Ternary Systems: Properties [70] K.S. Pitzer, Thermodynamics, 3rd ed. ed., McGraw-Hill, New York :, 1995. [71] S.R.S. Sastri, K.K. Rao, A simple method to predict surface tension of organic liquids, The Chemical Engineering Journal and The Biochemical Engineering Journal, 59 (1995) 181-186. [72] P.C. Simões, R. Eggers, P.T. Jaeger, Interfacial tension of edible oils in supercritical carbon dioxide, European Journal of Lipid Science and Technology, 102 (2000) 263-265. [73] J. Rao, D.J. McClements, Impact of lemon oil composition on formation and stability of model food and beverage emulsions, Food Chemistry, 134 (2012) 749-757. [74] C.M. Stafford, T.P. Russell, T.J. McCarthy, Expansion of polystyrene using supercritical carbon dioxide: effects of molecular weight, polydispersity, and low molecular weight components, Macromolecules, 32 (1999) 7610-7616. 186 Chapter 7 APPLICATIONS Chapter 7 shows the precipitation and foaming processes for the recycling of Polystyrene from its homogeneous solutions in terpene oils using CO 2 as antisolvent and blowing agent. Once the phases’ equilibrium was studied and the effect of CO 2 on the shift of physical properties was measured, the design of a foaming process was carried out. Initially, the feasibility of the processes was determined on the selection of the solubility parameters and the vanishing points of the solutions in CO2. Supercritical Antisolvent and Foaming Processes 188 Chapter 7 Based on: C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. De Lucas, M.T. García, Development of a strategy for the foaming of polystyrene dissolutions in scCO 2, Journal of Supercritical Fluids, 76 (2013) 126-134. C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. De Lucas, M.T. García, Foaming process from Polystyrene/p-Cymene solutions using CO2, Chemical Engineering & Technology, in press. C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. de Lucas, M.T. García, Preparation and characterization of polystyrene foams from limonene solutions, The Journal of Supercritical Fluids, 88 (2014) 92-104. Graphical abstract 189 Supercritical Antisolvent and Foaming Processes RESUMEN Las técnicas que emplean un antisolvente en estado supercrítico (SAS) se han convertido en una de las tecnologías más apropiadas para conseguir la precipitación de un polímero dando lugar a una morfología bien definida. El estudio del equilibrio termodinámico y las propiedades de las fases que forman parte del proceso es fundamental para seleccionar las condiciones de operación a las que llevar a cabo el proceso de precipitación. En primer lugar, se determinó la viabilidad del proceso de separación en base a la predicción de los parámetros de solubilidad a alta presión y la extracción del aceite terpénico, que contiene Poliestireno, con CO2. Sin embargo, cuando se plantea el proceso SAS en continuo, el Poliestireno no precipita de acuerdo a una morfología previamente seleccionada. Por este motivo, es necesario estudiar otros factores como la hidrodinámica del sistema, ya que es fundamental para lograr precipitar el polímero adecuadamente. Con el objetivo de preseleccionar el tamaño de las partículas de Poliestireno, se calcularon los coeficientes de difusión de las mezclas binarias para aplicarlos de acuerdo con la segunda ley de Fick. Así, se consiguieron las condiciones para precipitar el Poliestireno, a partir de su disolución en terpeno usando CO2 como antisolvente, dando lugar a partículas secas y con una estructura bien definida. Sin embargo, debido a la complejidad del proceso SAS, se considera más adecuada la espumación de residuos de Poliestireno para dar lugar a espumas microcelulares que presentan un alto valor añadido. En este apartado, inicialmente se determinó la viabilidad del proceso de espumación y a continuación, se llevó a cabo un diseño completo de experimentos. En un proceso de espumación, hay numerosos parámetros que permiten modificar las propiedades de las espumas. Por este motivo, se estudió el efecto de la presión, la temperatura, la concentración de la disolución, el tiempo de contacto y el tiempo de despresurización en el diámetro medio de las celdas y su distribución, así como en su densidad. Gracias a la presencia del disolvente terpénico, el proceso de espumación se puede desarrollar a condiciones de presión y temperatura moderadas. De acuerdo con el diseño factorial, se obtuvo que las condiciones más adecuadas para obtener espumas microcelulares de Poliestireno a partir de su disolución en Limoneno, fueron 90 bar, 30 ºC, 0.1 g PS/ml Limoneno, 240 minutos de tiempo de contacto y 30 minutos de despresurización. Por último, se analizaron las muestras del polímero recuperado para determinar la cantidad residual de disolvente en la espuma, la temperatura de transición vítrea y la temperatura de degradación, demostrando que el Poliestireno espumado contenía cantidades inferiores a un 5% de Limoneno, y en ningún caso hubo signos de degradación. 190 Chapter 7 ABSTRACT The Supercritical Antisolvent (SAS) techniques are shown as a very promising technology to achieve precipitation of polymer according to a well-defined morphology. The study of the thermodynamic properties of the phases involved in the process is crucial to select the most suitable conditions to perform the process. Initially, the feasibility of the separation process was studied according to the prediction of solubility parameters at high pressure and the extraction of the terpene from a polymer solution using CO2. Nevertheless, during continuous SAS process, Polystyrene did not precipitate according to previously selected morphology. By this reason, other factors such as the hydrodynamic of the system are crucial for that to succeed. Diffusion coefficients were calculated and Fick’s second law was applied with the aim of fixing the size particles. Finally, well defined and dried structures were obtained from the precipitation of Polystyrene solutions using CO 2. Nevertheless, because of the complexity of the process, the foaming of Polystyrene wastes was considered as a very valuable alternative to precipitate the polymer into a final product with high-added value. In this section, the feasibility of the foaming process was determined and next, a complete design of experiments was performed. During the foaming process, many parameters can be tuned to customize the foams. In this section, a design of experiment was proposed to determine the effect of pressure, temperature, concentration of the solution, contact time and vent time over the diameter of cells, its standard deviation and the cells density. The proposed foaming process can be simply performed at mild pressure and temperature thanks to the presence of the solvent. The results showed that the most suitable conditions to foam Polystyrene from Limonene solutions are 90 bar, 30 ºC, 0.1 g PS/ml Limonene, 240 minutes contacting and 30 minutes venting. Finally, the samples were characterised to determine the amount of residual solvent, their glass transition and degradation temperature checking that the foams presented around 5% of solvent traces but did not show any evidence of degradation. 191 Supercritical Antisolvent and Foaming Processes 192 Chapter 7 7.1. Supercritical Antisolvent Background Precipitation processes using a supercritical antisolvent allow solid microparticles with controlled size and morphology to be obtained from a raw product [1]. CO2 has been widely used as antisolvent for the preparation of polymeric microparticles due to its excellent properties, its mild supercritical conditions and the immiscibility of polymers into the gas [2-6]. An increasing number of processes in which CO2 can be used as antisolvent have been developed during the last decades. These processes have been also defined as precipitation with a compressed antisolvent (PCA) since the solute to be precipitated is dissolved in the organic solvent, and the solution is sprayed across a capillary into a compressed or supercritical fluid, which is miscible with the solvent, but behaves as an antisolvent for the solute [7, 8]. The general process is basically named Supercritical Antisolvent (SAS) although there are specific acronyms to denote the different techniques: GAS (gas antisolvent), PCA (precipitation by compressed antisolvent), ASES (aerosol solvent extraction system), SEDS (solution enhanced dispersion by supercritical fluids) or RESS (rapid expansion of supercritical solutions) are among the most common [9, 10]. The use of compressed or supercritical antisolvent for the precipitation of polymer presents several advantages over the conventional techniques (spray drying, mechanical comminution, recrystallization…): reduce the use of organic solvents, prevent thermal and chemical degradation of solute, decrease residual solvent concentration and control particle size and particle size distribution [11]. Depending on the process conditions, nozzle configuration and species involved, particles with different characteristics can be obtained [12-14]. Under appropriate operating conditions, high mass transfer occurs [6], the high diffusion rates and intense atomization can lead to rapid phase separation producing submicrometre particles [8]. The simultaneous diffusion of the antisolvent into the organic phase and the solubilisation of the organic solvent into the bulk antisolvent are enhanced due to the gas-like diffusivities of the CO2 and the density reduction of the expanded organic phase [15]. The way in which the liquid solution is dispersed in the CO2 at high pressure depends mainly on the pressure, when it is above the mixture critical point, mass transfer takes place by turbulent diffusion and mixing of the fluids is faster than droplet formation. Above the mixture critical pressure formation of droplets is not expected, due to the inexistence of surface tension between the solvent and the CO 2. On the other hand, when pressure is below the mixture critical points, the jet is disintegrated and mass transfer occurs by molecular diffusion through the solvent droplet interface [16, 17]. In this case, the precipitated particles usually exhibit different morphology and higher size than those precipitated above the mixture critical point, in which high supersaturation levels arise, affording smaller particles [11]. Either below or above the critical pressure of the mixture, carbon dioxide induces a liquid–liquid phase split in the polymer solution [16]. 193 Supercritical Antisolvent and Foaming Processes The phase equilibrium in the SAS process is well known from a theoretical point of view, but there is a serious lack of experimental data for the multicomponent systems of interest for this process [13]. However, the mechanism of particle formation is not very clear yet and in several cases, unsuccessful precipitations were reported, even when conditions were correctly selected on the basis of binary and ternary systems behaviour [18]. The reasons for this are that some authors considered only binary diagrams CO2/organic solvent due to the difficulties in the construction of ternary diagrams [19], the complexity of the process, in which phase behaviour, mass transfer, precipitation kinetics and hydrodynamics strongly interact with each other [16] needs also to be included. In this section, the precipitation of Polystyrene from its solution in terpenes using CO2 as antisolvent was studied. Initially, the technical feasibility of the separation process was studied theoretically from the solubility parameters point of view and experimentally from an extraction procedure. Next, the continuous process, where a solution of PS/terpene is fed to a vessel previously filled with high pressure CO 2, was performed. The mass transfer and the hydrodynamic of the system was studied in order to achieve some trends which relate the effect of pressure, temperature, concentration, flow or nozzle diameter on the morphology of the recovered polymer. 7.2. Supercritical Antisolvent Process 7.2.1. Solubility Parameters It has been demonstrated the power of solubility parameters as a tool for the prediction of solubility (Chapter 4). The methodology established by Hansen [20] was extended to be applied at high pressure by means of Equations of State. On a general basis, when the solubility parameters of a solute is within 4 MPa1/2 of that of the solvent, the solute will be pretty soluble in it. This rule can be applied to gaseous, liquid, crystalline and polymeric solutes [21]. In the antisolvent processes, the organic solvent should be soluble in CO 2, while the solid (PS) should not be. According to this fact, the solubility parameters () of CO2/PS and CO2/Limonene were calculated in vapour, liquid and solid phase in order to select the most suitable conditions to perform the antisolvent process. When the solubility parameters of liquid-vapour in the case of binary system CO2/Limonene are lower than 4 MPa1/2, while the solubility parameters of solidvapour in the binary system CO2/PS are higher, the antisolvent process could be accomplished. Figure 7.1 shows the solubility parameters of the defined systems as a function of pressure and temperature. The solubility parameters of the vapour phase are depicted with reddish tones, and the solubility parameters of the liquid or solid phases are shown as a bluish shade. 194 Chapter 7 5 50 4 40 3 30 2 20 1 10 0 25 30 35 Temperature (ºC) 40 0 10 20 30 40 50 Pressure (bar) 60 70 80 0 25 30 35 Temperature (ºC) 40 0 10 20 30 40 50 60 Pressure (bar) Figure 7.1 (Left) Effect of pressure and temperature on the solubility parameters of the vapour and liquid phase of the system CO 2/Limonene. (Right) Influence of pressure and temperature on the solubility parameters of the vapour and solid phase of the binary system CO2/PS. According to Figure 7.1 (left) is observed that CO 2 and Limonene are soluble in all the studied range of pressure (0-80 bar) and temperature (25-30 ºC) since the difference between the solubility parameters is lower than 4 MPa1/2. Regarding the solubility parameter of the vapour and liquid phase, the highest solubility of Limonene in CO2 is achieved in the range of pressure between 60 and 70 bar in the interval of temperature between 25 and 30 ºC because the solubility parameter in vapour and liquid phase are identical. On the other hand, Figure 7.1 (right) shows that independently of the pressure and temperature, the differences between the solubility parameters of solid and vapour phases is always higher than 4 MPa1/2 and the precipitation of PS should be occurred instantaneously. 7.2.2. Extraction process and recovery of solvent According to the explained previously, a first set of experiments was performed to remove the Limonene from the PS/Limonene solutions and obtain a solvent-free solid material. The first experiment was carried out to ensure that no Polystyrene/Limonene mixture dragging occurred during the extraction runs. In these experiments, the extractor was loaded with 10 g of mixture and CO 2 was fed at 77 bar, 40 ºC (= 250 Kg/m3) and Q = 2 L/min. The flow was selected previously, as the maximum value which avoids PS dragging according to the previous experience on oil extraction [22-24]. Figure 7.2 shows the Limonene concentration in CO 2 measured in the ternary mixture CO2/Limonene/PS. Limonene solubility was determined from the slope of the linear parts of extraction curves, drawn as grams limonene obtained versus grams CO2 used at the beginning of each extraction. The dash horizontal line 195 70 80 Supercritical Antisolvent and Foaming Processes represents the solubility of pure limonene in a binary system CO 2/Limonene (yLimonene = 0.00216), obtained from the literature [25]. 0.0024 Limonene solubility in CO2 0.0022 yLimonene 0.0020 0.0018 0.0016 0.0014 0.0012 0 1 2 3 4 5 6 7 8 Limonene in the mixture (g) Figure 7.2. Limonene concentration (in mole fraction) in the CO 2 exit stream on the amount of limonene in the PS/Limonene solution at 40 ºC and 77 bar. When the amount of Limonene in the solution increases the solubility of Limonene in CO2 rich phase increases due to the negative effect of the polymer concentration on the solubility of terpene in the vapour phase. Also it is observed that Limonene solubility in the ternary system is lower than in the binary one (CO 2/Limonene, dashed line), because the distribution coefficient decreases as a consequence of the presence of Polystyrene in the system [26]. These effects were explained in Chapter 5. The second experiment was performed to determine the extraction yield of the recovered Limonene. The selected conditions to run the experiment were shown above and it was carried out until Limonene extinction (Figure 7.3). 100 Yield (%) 80 60 40 20 0 0 1 2 3 4 5 6 7 8 Time (h) Figure 7.3. Yield evolution of the extraction of limonene. Conditions: P=77 bar; T=40 ºC; L= 10 g ; Q=2 L/min. 196 Chapter 7 Figure 7.3 shows that after 7 hours, the yield of the extraction of Limonene in the CO2 rich phase does not increase significantly, reaching a value of 95%. The remaining Limonene is not collected because the residence time of the solvent in the collector is not enough for the Limonene to condensate. These easy experiments were used to check that PS can be recycled from its solution in Limonene, using CO2 as antisolvent. The amount of Limonene in the polymer, its molecular weight distribution and the degradation temperature were studied in order to determine the quality of PS recovered after the recycling process. Table 7.1 shows the results from the thermal and calorimetric analysis and Figure 7.1. Table 7.1. Comparison between the properties of PS wastes and recycled. PS wastes PS recovered Glass transition temperature (ºC) 100.00 96.36 Degradation temperature (ºC) Polydispersity index 421.90 1.70 421.01 1.74 According to Table 7.1, the degradation of PS does not occur during the extraction since the characteristic of the original polymer were kept constant after the process. From the thermogravimetric analysis it is observed that the recycled PS is almost free of solvent. Therefore, considering the results obtained, it is possible to confirm that polymer degradation does not occur during the solvent removal using supercritical CO2. 7.2.3. Supercritical Antisolvent Process Regarding the promising results obtained for the recovery of PS wastes and Limonene by extraction with CO2, and the feasibility of the process according the solubility differences among the components, the Supercritical Antisolvent process was considered. The mentioned process was selected due to the growing interest that has arisen during the last decade for the development of micro- and nanoparticles. Once the feasibility of the separation was studied and with the aim of widen the study, the next plan of experiments was performed. For the design of a SAS process, the studied variables were pressure, temperature and concentration of PS in the solution, while the volume of solution feed, the flow and nozzle diameter were kept constant. Table 7.2 shows the range of the mentioned variables and the morphology of the recycled PS at 20 ºC. The results shown in Table 7.2 could be subdivided according to the working pressure (100, 150 and 200 bar). Regarding the pressure, any significance change in the morphology is observed in the range of study while the concentration is kept 197 Supercritical Antisolvent and Foaming Processes constant. Some authors have observed that the process conditions (mainly the pressure) have little influence on the particle size and morphology [27, 28]. Table 7.2. Summary of the operational conditions and results for the experiments performed. For all experiments: Temperature: 20 ºC; volume of solution: 25 ml; Qsolution: 0.5 mL/min; nozzle diameter: 0.0016 m. Run P (bar) CPS (g/ml) Morphology 1 100 0.004 Wet Microparticles 2 100 0.009 Wet agglomerated microparticles 3 100 0.050 Wet microparticles+ PS wet paste 4 100 0.100 Wet fibres 5 100 0.200 PS wet paste 6 100 0.250 PS wet membranes-like 7 100 0.500 Wet Fibres 8 150 0.045 Foamed PS 9 150 0.500 Foamed PS 10 200 0.050 Wet microparticles+ PS wet paste 11 200 0.200 PS wet paste On the other hand, the effect of concentration on the morphology of the Polystyrene shows that at constant pressure, when concentration of PS increases, the recovered polymer presents higher amounts of p-Cymene and consequently, the structure is agglomerated, wet and shows an amorphous appearance. Nevertheless, it is worth mentioning that when the concentration of PS in p-Cymene is higher than the maximum solubility (0.30 g PS/ml p-Cymene at 30 ºC, see references in Chapter 4), runs 7 and 9, a foamed PS is recovered. Dixon et al.[8] observed that spraying a PS/Toluene solutions into subcritical CO2, microspheres and microballoons were formed. However, when the concentration of polymer increased from 6 to 20%, microspheres were no hollow, but fully porous, with a continuous pore structure inside. This fact agrees with the results shown in this work. When temperature increased up to 40 ºC, while the volume of solution, its flow and the nozzle of the solution feed were kept constant, any precipitated polymer was recovered. Table 7.3 summarizes the most representative experiments performed at 40 ºC. 198 Chapter 7 Table 7.3. Summary of the operational conditions and results for the experiments performed. For all experiments: Temperature: 40 ºC; volume of solution: 25 ml; Qsolution: 0.5 mL/min; nozzle diameter: 0.0016 m. Run P (bar) C (g/ml) Morphology Final Concentration (g/ml) 12 100 0.045 Higher concentration solution 1.720 13 100 0.009 Higher concentration solution 1.890 14 100 0.200 Higher concentration solution 2.457 15 100 0.250 Higher concentration solution 2.886 16 150 0.045 Higher concentration solution 1.825 17 150 0.200 Higher concentration solution 2.474 18 200 0.050 Higher concentration solution 2.467 19 200 0.200 Higher concentration solution 2.346 According to Table 7.3, independently of the working pressure and concentration, at 40 ºC any solid PS was recovered. Nevertheless, an increase in the concentration of the solution (~ 2.25 g/ml) was observed as a consequence of the partial solubilisation of p-Cymene in CO2. The fact that any discrete particles are formed may be due to the agglomeration of droplets. The wet droplets collide with others from the wall of the vessel which makes more difficult their drying. Moreover, the flying time (from the nozzle to the bottom of the vessel) for agglomerated particles is too short for the drops to dry and they become sticky forming agglomerated droplets which finally make higher concentration solutions [28]. Also, Obrzut et al. [29] observed that at higher temperature, the jet breaks up as a gas-like plume without the formation of liquid droplets due to the dramatically decrease in the interfacial tension. In view of these results and considering the ternary equilibriums shown in Chapter 5, the ternary diagrams of the experiments shown above are depicted in Figures 7.4 to 7.6 in order to try to draw some general conclusions. 199 Supercritical Antisolvent and Foaming Processes e 1.0 0.2 0.8 0.4 0.6 0.6 e p-C en tyr ym lys en 0.0 Po Pressure: 100 bar; T: 20 ºC Run Solution SAS 1 2 3 4 5 6 7 0.4 0.8 0.2 1.0 0.0 0.0 0.2 0.4 0.6 0.8 1.0 CO2 Figure 7.4. Ternary diagram comparing mass-transfer pathways for precipitation of PS/pCymene solutions (filled markers) with compressed CO 2 as antisolvent at 100 bar and 20 ºC. The composition of the ternary mixtures inside the vessel has been depicted as empty symbols. Pressure: 150 bar; T: 20 ºC Run Solution SAS 8 9 0.0 1.0 0.2 en 0.4 0.6 lys en p-C tyr ym Po e 0.8 e 0.6 0.4 0.8 0.2 1.0 0.0 0.0 0.2 0.4 0.6 0.8 1.0 CO2 Figure 7.5. Ternary diagram comparing mass-transfer pathways for precipitation of PS/pCymene solutions (filled markers) with compressed CO 2 as antisolvent at 150 bar and 20 ºC. The composition of the ternary mixtures inside the vessel has been depicted as empty symbols. 200 Chapter 7 Pressure: 200 bar; T: 20 ºC Run Solution SAS 10 11 0.0 1.0 0.2 0.8 ym p-C en tyr 0.6 lys en e Po 0.4 e 0.6 0.4 0.8 0.2 1.0 0.0 0.0 0.2 0.4 0.6 0.8 1.0 CO2 Figure 7.6. Ternary diagram comparing mass-transfer pathways for precipitation of PS/pCymene solutions (filled markers) with compressed CO 2 as antisolvent at 200 bar and 20 ºC. The composition of the ternary mixtures inside the vessel has been depicted as empty symbols. From Figures 7.4-6, it is observed that an increase in the initial concentration of PS in the solutions promotes an increase in the slope of the tie lines, as it was previously described in Chapter 5. However, the increase of pressure does not produce any significantly shift on the position of the markers in the ternary diagrams. From the comparison of Figures 7.4-6 with the Figure 5.9, it is checked that during the SAS process, the composition of all the ternary mixtures inside the vessel is located in the two-phase regions. By this reason, two phases should be recovered, a polymer rich phase and a vapour rich phase which contains CO 2 and pCymene. The amount of solvent of the recovered polymer depends on the position of the tie line. Next, the ternary diagram of the experiments performed at 100 bar and 40 ºC is shown (Figure 7.7). 201 Supercritical Antisolvent and Foaming Processes Pressure: 100 bar; T: 40 ºC Run Solution SAS 1 2 3 4 0.0 1.0 0.2 en 0.4 0.6 lys en p-C ty r ym Po e 0.8 e 0.6 0.4 0.8 0.2 1.0 0.0 0.0 0.2 0.4 0.6 0.8 1.0 CO2 Figure 7.7. Ternary diagram comparing mass-transfer pathways for precipitation of PS/pCymene solutions (filled markers) with compressed CO 2 as antisolvent at 100 bar and 40 ºC. The composition of the ternary mixtures inside the vessel has been depicted as empty symbols. From the comparison of Figure 7.7 obtained at 40 ºC and Figure 7.4 at 20 ºC, both at 100 bar, any remarkable difference is observed, although at higher temperature any solid is recovered as it was also shown in Table 7.3. Considering these results it could be thought that the feasibility of the SAS process requires not only knowledge of the phase equilibrium ternary systems, but also the hydrodynamics of the system [17], the nozzle design and the ways of mass transfer [30]. When the solution is fed to the high pressure vessel different regimes could be produced: the Rayleigh breakup, the sine wave breakup and the atomization. On a regular basis, for the precipitation of homogeneous microparticles, atomization is preferred, but when the working pressure is located above the mixture critical point, it is not possible to distinguish droplets since there is not interfaces between the solution and the dense CO2 [17, 31]. Taking into account the importance of the hydrodynamics of the system, a relation between the experimental morphology of the PS recovered and the regimes was attempted. Czerwonatis and Eggers defined two new dimensionless numbers, Z* and Z**: [7.1] 202 [7.2] Chapter 7 where l is the viscosity of the liquid, is the velocity of the liquid jet, is the interfacial tension and g and l represent the density of the gas and liquid phase, respectively. By representing the Z* and Z** numbers vs Reynolds (Re) and Morton (Mo) dimensionless numbers of the liquid phase it would be possible to determine the limits of the working conditions. The mentioned dimensionless numbers are defined as: [7.3] Re expresses the ratio of fluid inertial and viscous forces, where v is the mean velocity of the fluid, D is the hydraulic diameter, is the density of the fluid, is the dynamic viscosity. [7.4] Mo represents the characteristics of the shape of bubbles and/or drops moving in a surrounding fluid; where g is the acceleration of gravity, c is the viscosity of the surrounding fluid, c is the density of the continuous phase, is the difference in density of the phases and is the surface tension. The density of the solutions (was calculated according to the equations developed by Ref. for the compressed liquid mixture [32]. The dynamic viscosity ( was obtained from the experimental values of viscosity shown in Chapter 6 and the velocity of the fluid. The viscosity of CO2 (c) at high pressure was calculated following the Lucas method for compressed gases at high pressure [33], the density of the surrounding fluid (c) was obtained from the Bender equation [34] and finally, the surface tension ( was obtained experimentally and data were shown in Chapter 6. The physical properties obtained for the calculations at the different temperature are listed in Table 7.4. Table 7.4. Physical properties for the system CO 2/p-Cymene/PS at 20, 30 and 40 ºC. Volume of solution: 25 ml; Qsolution: 0.5 mL/min; nozzle diameter: 0.0016 m. Run v l P T g G·106 (bar) (ºC) (m/s) (kg/m3) (kg/m3) (kg/m·s) G·10-6 (m2/s) 1 100 20 0.004 857.159 846.121 84.420 0.099 20 100 30 0.004 771.500 846.489 66.160 0.086 12 100 40 0.004 622.641 846.833 47.840 0.076 Run v P T L L·106 (bar) (ºC) (m/s) (kg/m·s) (m2/s) t=0 (N/m) vcrit (m/s) 1 100 20 0.004 0.005 5.909 0.003 1.301 20 100 30 0.004 0.003 3.544 0.003 1.371 12 100 40 0.004 0.002 2.362 0.001 0.922 203 Supercritical Antisolvent and Foaming Processes For the study of the hydrodynamic regimes a wide set of experiments was considered, in which the solution flow was varied from 0.2 to 3 mL/min. The morphology of the recovered PS was classified in three different categories: foams, wet microparticles or higher concentration solutions as it was shown in Tables 7.2 and 7.3. The study of Z* and Z** numbers vs Re and Mo at different operating conditions are shown in Figure 7.8. 0.010 0.100 b) Z** vs Re Foams Wet microparticles 0.075 Solutions 0.008 Z** Z* 0.006 0.004 0.050 * a) Z vs Re Foams Wet microparticles Solutions 0.002 0.000 0 1 2 3 0.025 0.000 4 5 6 0 Re 0.020 0.15 c) Z* vs Mo Foams Wet microparticles 0.015 Solutions 1 2 3 4 5 6 Re d) Z** vs Mo Foams Wet microparticles Solutions Z** Z* 0.10 0.010 0.05 0.005 0.0 2.0x10 -5 4.0x10 Mo -5 6.0x10 -5 0.00 0.0 2.0x10 -5 4.0x10 -5 6.0x10 -5 Mo Figure 7.8. Relation between the hydrodynamic conditions and the morphology of the recovered PS from a PS/p-Cymene solution in pressurized CO2. According to Figure 7.8 several trends could be observed but in general, the low values of Re and Mo dimensionless numbers should be highlighted. Since the Reynolds number describes the inertial force to friction force ratio, the droplet size should decrease with increasing Re number. But some authors observed that the smallest particles were produced in the subcritical region, where the Re number was either very small [28]. After considering the results shown, Rayleigh breakup was the only regime achieved due to the low values of Re and Mo, although the critical points of the mixtures were reached (see Chapter 6). From Figures 7.8 a and b (Z* vs Re and Z** vs Re, respectively), at higher values of Re and Z* or Z**, any dried structure was observed. On the other hand, from Figures 7.8 c and d (Z* vs Mo and Z** vs Mo, respectively) three zones are identified. Higher concentration solutions are observed when Mo is around 510-5, Z*=0.075 and Z**=0.05, while when Z* and Z** increases, PS/p-Cymene solutions precipitate in the form of wet microparticles 204 Chapter 7 and when Z* and Z** decreases, PS/p-Cymene are foamed. Agglomerated particles could lead into foams; this fact could be explained by the increase of CO2 sorption in the polymeric solution (see Chapter 5) and the existence of significant mechanisms that stabilize the liquid jet. It means that although equilibrium interfacial tension is zero (Table 7.4.), a certain interface exist between the solution jet and the surrounding gas [17]. Attending to the results, changes in the fed flow (velocity) does not have any significant effect on the morphology of the precipitated PS [28], although faster solution flow rates should produce spherical particles with increasing diameters while fibres are generally attributed to the precipitation before the jet break up [1]. The morphology of the precipitate obtained at pressure above the mixture critical point should have microparticle shape, since the prerequisites for successful antisolvent process are the complete miscibility between the organic solvent and the antisolvent and the insolubility of the polymer in the antisolvent [35]. However, dried microparticles or fibres were not precipitated from the PS/p-Cymene solutions. With the aim of study the effect of solvent into the precipitated PS in order to achieve microparticles, the system CO 2/Limonene/PS was studied following the procedure described by the system CO2/p-Cymene/PS. According to Figure 7.9 similar trends to those observed in the ternary system CO2/p-Cymene/PS are shown. In the case of PS/Limonene solutions, Re dimensionless number increases, but Mo dimensionless numbers decreases one order. The decrease of Mo implies little deformation of the droplets that acquire a nearly spherical shape instead of ellipsoidal [36] which is result of a higher relative strength of interfacial forces versus viscosity. Viscosity values of PS/Limonene and PS/p-Cymene were similar, but when the concentration of PS in Limonene increases the interfacial tension of the mixture increases reaching values higher than in the case of PS/p-Cymene solutions (Figure 6.13). By this reason, the interfacial forces prevail over the viscosity which promotes the formation of spherical droplets in PS/Limonene solutions. However, due to the low value of Re and Mo, Rayleigh breakup was also the regime achieved. 205 Supercritical Antisolvent and Foaming Processes 0.3 a) Z* vs Re Foams Wet microparticles Solutions 0.10 0.08 b) Z** vs Re Foams Wet microparticles Solutions 0.2 Z** Z* 0.06 0.04 0.1 0.02 0.00 0.0 0 2 4 6 12 14 16 18 20 0 0.15 c) Z* vs Mo Foams Wet microparticles Solutions 2 4 6 8 10 12 14 16 18 20 Re d) Z** vs Mo Foams Wet microparticles Solutions 0.10 0.015 Z** Z* 0.020 10 Re 0.030 0.025 8 0.010 0.05 0.005 0.000 0.0 0.00 5.0x10 -7 1.0x10 Mo -6 1.5x10 -6 2.0x10 -6 1.0x10 -6 2.0x10 -6 3.0x10 -6 4.0x10 -6 Mo Figure 7.9. Relation between the hydrodynamic conditions and the morphology of the recovered PS from a PS/Limonene solution in pressurized CO 2. Figures 7.9 a and b show the trend line which separate the precipitation of PS as a higher concentration solution or as foams. Wet microparticles can not be summarized according to any specific trend line. Nevertheless, in Figures 7.9 c and d clear trends are shown for the precipitation of wet microparticles or foams, although the formation of higher concentration solutions is performed at low values of Mo (10-6) and Z* around 0.005 or Z** 0.025. According to the literature, decreasing the initial affinity of the solute for a solvent (PS presents lower solubility in Limonene than in p-Cymene, Chapter 4) should precipitate larger particles because they stay suspended in the solution longer and continue to grow [37]. Nevertheless, from the comparison of Figure 7.9 and 7. 8 this fact was not observed maybe due to the similarities between both solvents. Considering the difficulties to the precipitation of PS into homogeneous microparticles shape, several authors has determined that the fast disappearance of the surface tension of the injected solution is the cause of microparticles nucleation and growth. When the liquid and the supercritical CO 2 are contacted, mutual diffusion leads to a progressive decrease of the dynamic surface tension [38] and consequently, the organic solvent is solubilised into the CO 2 while the polymer is precipitated. By this reason and taking into account that mass transfer is driven by 206 Chapter 7 diffusion, the diffusivity coefficients of the binary systems CO2/p-Cymene, pCymene/CO2 and CO2/Polystyrene were calculated as a function of pressure and temperature. It should be outlined that the diffusion coefficients of CO 2/polymeric solutions were not calculated because there was any experimental value to check the prediction. For the calculation of the diffusion coefficients, several correlation were found in the literature: Hayduck-Minhas [32], He-Yu [39], Wilke-Chang [32], Reddy-Doraiswamy [40], Catchpole-King [40], Scheibel [41] and Lusis-Ractliff [40]. All of them were tested initially against some experimental diffusion coefficients reported in the literature for organic solvents [40, 42] and Hayduck-Minhas correlation was the selected to calculate the diffusion coefficients of CO2/p-Cymene as follows: [7.5] [7.6] where Dab is the diffusion coefficient of substance a in b, T is the temperature, a is the viscosity of substance a and Vb is the molar volume of substance b. The viscosity of CO2 was calculated following the Lucas method for compressed gases at high pressure [33] the viscosity of p-Cymene solutions was determined experimentally and results were shown in Chapter 6 and the viscosity of PS was obtained experimentally by the use of a rheometer (Chapter 4) . On the other hand, the molar volume of CO2 and PS was calculated from the Sánchez-Lacombe EoS but the molar volume of p-Cymene was obtained from the Peng-Robinson EoS (see Chapter 4). This analysis does not consider the effect of the polymer on the increase in viscosity of p-Cymene since the binary systems were studied individually. Figure 7.10 shows the diffusion coefficients of the binary system CO 2/p-Cymene and p-Cymene/CO2. x 10 -5 CO2/p-Cymene 5 Diffusivity Coefficient (m2/s) 4.5 4 3.5 3 2.5 2 1.5 1 0.5 20 25 200 150 30 100 35 Temperature (ºC) p-Cymene/CO2 50 40 0 Pressure (bar) Figure 7.10. Diffusion coefficients of the binary systems CO 2/p-Cymene and p-Cymene/CO2. 207 Supercritical Antisolvent and Foaming Processes Figure 7.10 shows that diffusion of CO2 inside the droplet and diffusion of p-Cymene from the droplet to the surrounding CO 2 is balanced at the lowest pressure, while when pressure increases the diffusion coefficient of CO 2 inside p-Cymene increases and its sorption is enhanced but the solubilisation of p-Cymene is not. According to Figure 7.10, the antisolvent process should be performed at atmospheric pressure, nevertheless, the solubility of p-Cymene in CO2 at low pressure is negligible (see Chapters 4 and 5). The effect of pressure and temperature on the diffusion coefficients of CO 2 into Polystyrene was predicted according to He and Yu [43] correlation for supercritical fluids. The calculated data were checked experimentally (Chapter 4) and results are shown in Figure 7.11. Figure 7.11 shows that when pressure increases, the diffusion coefficient of CO 2 into Polystyrene increases, as well as when temperature increases. Although at low pressure, the effect of temperature is not significantly. In this case, the diffusion coefficient of Polystyrene into CO2 was not calculated since its solubility is negligible [44, 45]. x 10 -10 Diffusivity Coefficient (m2/s) 1.8 1.6 1.4 1.2 20 25 200 150 30 100 35 Temperature (ºC) 50 40 Pressure (bar) Figure 7.11. Diffusion coefficients of CO2 into Polystyrene. From the comparison of Figures 7.10 and 7.11, it is observed that solubility of CO2 is enhanced in p-Cymene solutions, while the presence of the polymer would decrease it because the DCO2/PS presents 5 orders of magnitude lower. This fact could explain the results obtained in Chapter 5 (Figure 5.11), when the concentration of Polystyrene increased, the solubility of p-Cymene in CO2 decreased. Finally, the suitability of the selected ternary system to carry out the antisolvent process was checked by the several differences between the diffusion coefficients. 208 Chapter 7 Once the diffusion coefficients are obtained, the concentration of each compound as a function of time and position could be determined using the Fick’s Second Law of diffusion [46]. [7.7] where C is the concentration, t is the time and x is the studied position. The solutions to Fick’s second law depend on the boundary conditions imposed by the particular case; in this case, a spherical droplet of p-Cymene with radius r was supposed. The radius of the droplet can be calculated from the nozzle diameter because when a liquid is forced through a nozzle, at the nozzle exit it forms a continuous cylindrical jet with a diameter equal to that of the nozzle [47]. Attending to the spherical geometry of the droplet, expression [7.7] becomes: [7.8] If the substitution u=Cr is made and the boundary conditions shown below [7.97.12] are applied, the Fick’s second law equation is expressed as equation [7.12]. u=0, r=0, t>0 [7.9] u=aC0, r=a, t>0 [7.10] u=rf(r), 0< r < a, t=0 [7.11] [7.12] where C is the concentration of the surface of the sphere, C 0 is the initial bulk concentration and erf refers to the Gaussian error function. According to the expression [7.12] and considering the diffusion coefficients shown in Figure 7.10, the concentration of p-Cymene and CO2 in relative terms along time in a droplet of radius 7.94·10-4 m (experimental nozzle diameter) is shown in Figure 7.12. 209 0 10 10 20 Time (s) 0 CO2 inside the droplet Temperature: 20 ºC 30 P: 1 bar P: 100 bar P: 200 bar 0 p-Cymene outside the droplet Temperature: 20 ºC P: 1 bar 30 P: 100 bar P: 200 bar 40 0 10 10 20 CO2 inside the droplet Temperature: 30 ºC 30 P: 1 bar P: 100 bar P: 200 bar 0 40 p-Cymene outside the droplet Temperature: 30 ºC P: 1 bar 30 P: 100 bar P: 200 bar 10 Time (s) 20 20 0 40 10 Time (s) 20 40 Time (s) Time (s) Time (s) Supercritical Antisolvent and Foaming Processes CO2 inside the droplet Temperature: 40 ºC 30 P: 1 bar P: 100 bar P: 200 bar 40 0.0 0.2 0.4 0.6 Concentration 0.8 1.0 20 p-Cymene outside the droplet Temperature: 40 ºC P: 1 bar 30 P: 100 bar P: 200 bar 40 0.0 0.2 0.4 0.6 Concentration 0.8 1.0 Figure 7.12. Influence of pressure and temperature on the relative concentration of CO2 inside the droplet (left figures) and p-Cymene outside the droplet (right figures) along the time p-Cymene/CO2 system. According to Figure 7.12, the solubility of CO2 inside the droplet increases dramatically up to 5 seconds, independently of working pressure and temperature. But as time goes over 5 seconds, the droplet saturation is achieved since the concentration of CO2 inside the droplet remains almost constant in a range of concentration between 0.91 and 0.96 (in relative terms of concentration), depending on the operating pressure and temperature. On the other hand, the effect of pressure and temperature affects on the solubility of p-Cymene in the CO2 rich phase (right figures) significantly. When pressure is above the mixture critical point (> 70 bar), the solubility of the solvent outside the droplet remains almost constant independently of the temperature. Nevertheless, higher temperatures promote a faster diffusion of p-Cymene from the droplet to the surrounding media. Regarding Figure 7.12, dried PS could not be recovered at the studied working conditions since the concentration of p-Cymene outside the droplet does not reach 1. This fact was observed experimentally in Table 7.2 and 7.4. 210 Chapter 7 0 0 10 10 20 Time (s) Time (s) The initial size of the droplet depends upon the nozzle diameter and initial velocity of the droplet in addition to the physical properties of the organic solvent, such as, density, surface tension, and viscosity [48]. Considering the results shown a decrease in the diameter of the droplet could improve the diffusivity of p-Cymene outside the droplet, and consequently, dried PS could be achieved. By this reason, new simulations were performed to study the influence of the nozzle diameter on the concentration of p-Cymene in the vapour rich phase along the time. Figure 7.13 represents the evolution of CO2 and p-Cymene concentration in liquid and vapour phase, respectively through a 500 nm nozzle. CO2 inside the droplet Temperature: 20 ºC P: 1 bar P: 100 bar P: 200 bar 0 40 30 Time (s) Time (s) 10 CO2 inside the droplet Temperature: 30 ºC 30 P: 1 bar P: 100 bar P: 200 bar 0 40 p-Cymene outside the droplet Temperature: 30 ºC P: 1 bar 30 P: 100 bar P: 200 bar 10 Time (s) Time (s) 20 0 40 10 20 p-Cymene outside the droplet Temperature: 20 ºC P: 1 bar 30 P: 100 bar P: 200 bar 40 0 10 20 20 CO2 inside the droplet Temperature: 40 ºC 30 P: 1 bar P: 100 bar P: 200 bar 40 0.0 0.5 0.9990 0.9995 Concentration 1.0000 20 p-Cymene outside the droplet Temperature: 40 ºC P: 1 bar 30 P: 100 bar P: 200 bar 40 0.0 0.5 0.9990 0.9995 1.0000 Concentration Figure 7.13. Influence of pressure and temperature on the relative concentration of CO2 inside the droplet (left figures) and p-Cymene outside the droplet (right figures) along the time p-Cymene/CO2 system through a 500 nm nozzle. Figure 7.13 shows that a decrease in the nozzle size produces smaller drops that are dried faster since the concentration of p-Cymene in the surrounding media reaches saturation at times shorter than 1 second when the pressure is above the mixture critical point. As it was observed in Figure 7.12, the effect of pressure affects 211 Supercritical Antisolvent and Foaming Processes significantly on the diffusion of p-Cymene outside the droplet but reducing the nozzle diameter up to 500 nm, saturation of CO2 can be easily achieved from 100 bar. Thanks to these encouraging results and with the aim of generating small, homogeneous and fine droplets with a high surface area, the nozzle was replaced by a smaller one (500 nm). Rantakylä determined the efficiency of an antisolvent process by the droplets size produced in the nozzle and by the way in which the gaseous phase mixes with the droplets [28]. In this cases, when the nozzle was smaller, dried fibres or microparticles were produced. Figure 7.14 and 7.15 show some of the most representative microphotographs of the recovered Polystyrene from its precipitation in CO2 through a nozzle of 500 nm. (a) (b) Figure 7.14. Microphotographs of the precipitated PS using CO 2 as antisolvent at 20 ºC; Volume of solution: 10 mL. a) P: 100 bar, C: 0.05 gPS/ml p-Cymene, Qsolution: 0.5 mL/min; b) P: 200 bar, C: 0.05 gPS/ml p-Cymene, Qsolution: 0.5 mL/min. (a) (b) Figure 7.15. a) P: 100 bar, C: 0.05 gPS/ml Limonene, Qsolution: 1.8 mL/min; b) P: 200 bar, C: 0.05 gPS/ml Limonene, Qsolution: 1.9 mL/min. 212 Chapter 7 From Figures 7.14 and 7.15, it is observed that well defined and dried structures were obtained independently of the operating conditions and the solvent. The average ratio of microparticles was 250 m and the average size of the fibres diameter was 20-60 m. An increase in pressure, from 100 to 200 bar does not produced any significant change in the morphology of the precipitated polymer. However, an increase in the liquid flow promotes the coalescence of droplets which leads to fibres (Figure 7.15). This fact could be the issue that limits the volume of the solution and the concentration of polymer due to the clogging of the nozzle [28]. According to the results, the precipitation of PS from its solution in Limonene or pCymene using CO2 as antisolvent is feasibly. Nevertheless, important efforts should be carried out to achieve well defined polymeric structures due to the complexity of the process and the numerous variables involved. 213 Supercritical Antisolvent and Foaming Processes 7.3. Foaming background Microcellular foams are defined as foams with average cell sizes of less than 10 microns and cell densities greater than 109 cells/cm3.They are the target products since they typically exhibit high impact strength, toughness and thermal stability, as well as low dielectric constant and thermal conductivity [49]. The use of Supercritical Fluids (SCFs) as physical blowing agents has being developed for the production of polymer foams because they can create highly functional materials with new features and dramatically improve conventional manufacturing methods, especially from an environmental point of view. Currently, the focus is on carbon dioxide due to its relative ease of handling and more favourable interaction with polymers compared to other inert gas candidates like nitrogen [50]. The key advantages related with the usage of scCO2 to foam are due to the easily removal from the final product, its tunable solvent power and its capability to modify the polymer properties as a result of its sorption [51]. Having diffusion coefficients higher than that of liquids, viscosities close to that of gases and low surface tension, scCO2 provides better penetration and complete wetting of the substrates which is advantageous for impregnation or extraction applications as compared to conventional solvents [52]. Also, scCO2 has received an increasing attention with respect to traditional blowing agents, i.e. chlorofluorocarbons (CFCs), hydrochlorofluorocarbons (HCFCs) or hydrocarbons, due to its low cost and mild impact on the environment. Several authors have demonstrated that CO 2 can be used to foam amorphous materials such as poly(methylmethacrylate), polystyrene, polycarbonate, and poly(ethyleneterephthalate). Table 7.5 shows a good sample of the available literature data [49, 51, 53-69] for the foaming of polymers or polymer blends using supercritical CO2 and the working conditions used to carry out the process: pressure, temperature, contact time and depressurization rate ranges. Generally high pressure and temperature are required to foam the polymers but also there are other parameters playing an important role during the process. By this reason, polymeric foams can be customized and many polymers are screened for their foamability and resulting foam structure, which determines to a great extent the properties of the foam [68]. 214 90-160 300 69-210.7 100-200 150-200 86-140 180-380 80-230 140-610 100-250 103-276 230 70-280 78-200 250-450 120-350 150-190 Poly(ethylene terephthalate) isotactic Polypropylene Polystyrene Polystyrene Polystyrene Polystyrene Polystyrene and Poly(D,L-lactic acid) Polystyrene and cellulose acetate Poly(l-lactic acid) Poly(lactic acid-co-glycolic acid) Poly(D-lactide) and Poly(D-lactide-co-glycolide) Poly(D,L-lactide) and Poly (D,L-lactide-co-glycolide) Poly(e-caprolactone-co-lactide) Poly(e-caprolactone) Poly(p-dioxanone) Poly(heteroarylenes) Polypropylene, Polystyrene and Poly(styrene-butadiene-styrene) Polystyrene, Polymethylmethacrylate, Poly (styrene-co-butadiene-comethylmethacrylate) and Poly (methylmethacrylate-co-butylacrylate-comethylmethacrylate) 300 80-160 25-80 160-190 40-100 80-100 40-50 60-80 90 35-100 35-85 60-180 55-125 35-55 60-110 100 50-110 60-180 200 180-260 280 0.03-1 MPa/s 120 min 600 min 0.01 MPa/s 3-120 sec 60-120 min [62] [61] [60] [51] [59] [69] [68] [67] [66] [65] [64] 0.006-0.013 MPa/s [63] 0.3-0.4 MPa/s 0.05 -0.20 MPa/s 20 seconds 120 min 1-40 min 20 min 300 min 60-240 min 0.1-1.6 MPa/s 1 MPa/s 240 min 240 min [57] 80-400 MPa/s 0-60min [58] [49] 100-400 MPa/s 240 min [56] [55] [54] [53] 0.91-1.17 MPa/s 180-180 MPa/s 24 h 20-25 min ~ 60 min 30-50 min Pressure Temperature Contact time Depressurization Refs range (bar) range (ºC) range rate or time Poly(ethylene terephtalate) Polymer Table 7.5. Literature review of foamed polymers. Range of pressure, temperature, contact time and depressurization used. Chapter 7 215 Supercritical Antisolvent and Foaming Processes The process to prepare foams using CO 2 as foaming agent can be divided into three steps: (1) sorption of CO2 until saturation in the plastic; (2) nucleation of foam bubbles, it is need induce a phase separation by a thermodynamic instability (either a temperature increase or a pressure decrease); and (3) growth of foam’s cells. The main steps are schematically depicted in Figure 7.16. Figure 7.16. Schematic diagram of the steps during a typical foaming process: sorption, saturation, foaming and cell growth. The sorption of CO 2 into the polymer (continuous line, left axis) is the responsible of the decrease of the glass transition temperature (dashed line, right axis). As shown in Figure 7.16, the foaming process starts by melting the polymer and adding CO2. The type of polymer, together with the applied pressure and temperature determine to a large extent the amount of CO 2 that can be dissolved into. During the sorption step, the gas is dissolved inside the polymer inducing swelling, increasing its free volume and segment mobility, causing its plasticization which involves the modification of the mechanical and physical properties of the polymers. The absorbed gas behaves like a “lubricant” of the polymeric chains which are easier slid and consequently, the glass transition temperature (Tg), melting temperature (Tm), viscosity () and surface tension () are decreased. The nucleation stage occurs instantaneously at lower Tg when the polymer matrix becomes less rigid. Cell growth will stop once the polymer matrix returns to the glassy state, either due to a decrease in temperature or a decrease in the CO 2 concentration in the polymer and the polymer is no longer plasticized [70]. 216 Chapter 7 During the depressurization stage, run at constant temperature, the pressure is reduced, the CO2 phase changes from supercritical to gas and nucleation and growth of gas bubbles occurs in the rubbery polymer generating pores. However not only depressurization rate should be control, but also the cooling rate, because it affects the polymer viscosity (Figure 7.16) [63]. Important efforts have been carried out to determine the real influence of the experimental parameters on the final foam structure and as a consequence on their properties. In general, increasing pressure, the pore diameter and the bulk density decreases, while the cell density increases. At higher pressure, more fluid is dissolved into the polymer matrix causing a more pronounced plasticization and viscosity reduction. Nevertheless, the pore cell increases while the bulk foam density and cell population decreases when temperature is increased [59]. In this section, p-Cymene and Limonene were used as solvents for the PS since in Chapter 4 and Chapter 5, the high solubility of PS in terpenes and of terpenes in CO2 was checked. Furthermore, they allow to run the foaming process feasibly at mild working temperature. Moreover, the use of binary fluid mixtures of carbon dioxide and an organic solvent provide flexibilities towards control of pore sizes and connectivity in forming porous matrices [60]. CO2 is behaved as nonsolvent of the polymer and the mass transfer in the systems occurs in two ways, the Limonene is solubilised from the bulk polymeric phase at the same time as CO 2 is absorbed in the liquid phase. The mass transfer during the foaming is mainly governed by the diffusion of the CO2 into the polymeric rich phase and it has been checked that is enhanced because of the presence of the solvent which promotes the relaxation of the polymeric chains. According to these facts, the first aim consisted on the verification of the feasibility of the foaming process in the surroundings of the vanishing point, where the interfacial tension is vanished. Next, the study of the foaming process of a PS/pCymene solution using CO2 as antisolvent was carried out. Finally, the third purpose was the preparation, characterization and optimization of the foams to obtain high cells density and low pore cells by mean of a process which required low energy consumption and does not affect the polymeric chain structure because of the low operation temperature. 7.4. Feasibility of the Foaming of Polystyrene/cosolvent system with CO2 In this section, the feasibility of the foaming process of PS/terpene oils solution was studied. The selection of the operating limits was performed according to the glass transition temperature, viscosity and interfacial tension of the mixtures (Chapter 6). Generally low interfacial tension (IFT) is desired to increase the nucleation rate and produce small and uniform cells [71-73]. In Chapter 6 was shown that IFT of PS/p- 217 Supercritical Antisolvent and Foaming Processes Cymene and PS/Limonene solutions in presence of CO2, vanishes around 70-80 bar, depending on the temperature and initial concentration. These results indicate that Polystyrene should be able to foam at temperatures only slightly above room temperature, because IFT at mild working conditions practically vanishes and presents a pronounced decreasing rate. Also, the effect of CO2 on the glass transition temperature of PS and its solutions is crucial to limit the working range of the temperature during the foaming process. This parameter is key since the saturation of CO2 into polymer can occur either in a glassy or rubbery state, but nuclei growth can mainly occur in the rubbery zone or near the Tg of an amorphous polymer. Considering that the growth of nucleis in foams are achieved around Tg and low IFT are recommended to get homogeneous foams, the selected working conditions were 70-90 bar and 30-40 ºC. Taking under consideration the mentioned facts, a set of PS foams were produced at different temperatures, pressures and initial concentrations, while contact and depressurization time were kept constant (at 240 and 1.5 minutes, respectively). The working conditions were selected based on Reverchon’s studies [51]. On this way, the temperature values assayed was 30 and 40ºC, pressures 70 and 90 bar and contact time was set at 240 minutes in all cases to assure homogeneous microcellular structured. The structure and cell size of the processed PS foams analyzed by optical microscopy are shown in Figure 7.17 (from PS/p-Cymene solutions) and Figure 7.18 (from PS/Limonene solutions). Figure 7.17. Micrographs with 50x magnification of the foams obtained as a function of the initial concentration, temperature, pressure and solvent used. Cymene p: 70 bar p: 90 bar 218 T:30ºC C0=0.05 g/ml C0=0.20 g/ml T:40ºC C0=0.05 g/ml C0=0.20 g/ml Chapter 7 Figure 7.18. Micrographs with 50x magnification of the foams obtained as a function of the initial concentration, temperature, pressure and solvent used. Limonene T:30ºC C0=0.05 g/ml C0=0.20 g/ml T:40ºC C0=0.05 g/ml C0=0.20 g/ml p: 70 bar p: 90 bar As it is observed in Figure 7.17 and 7.18, foams were not produced at 70 bar and 40 ºC, independently of the concentration and solvent used. According to Tables 7.6 and 7.7 (runs 5 and 6) it is observed that at 70 bar and 40 ºC, the CO 2 mass fraction in the system is the lowest, and probably it is not sufficient to thoroughly solubilise the p-Cymene nor the Limonene and PS can not be foamed. Table 7.17. Mass fraction composition of the system adopted for PS foam synthesis from pCymene solutions. Subscripts: 1: CO2; 2: p-Cymene; 3: PS. Run P (bar) T (ºC) C (g/ml) w1 w2 w3 1 70 30 0.05 0.9114 0.0837 0.0049 2 70 30 0.20 0.8982 0.0825 0.0193 3 90 30 0.05 0.9664 0.0318 0.0019 4 90 30 0.20 0.9610 0.0316 0.0074 5 70 40 0.05 0.8843 0.1094 0.0064 6 70 40 0.20 0.8677 0.1073 0.0250 7 90 40 0.05 0.9493 0.0479 0.0028 8 90 40 0.20 0.9415 0.0475 0.0111 219 Supercritical Antisolvent and Foaming Processes Table 7.18. Mass fraction composition of the system adopted for PS foam synthesis from Limonene solutions. Subscripts: 1: CO 2; 2: Limonene; 3: PS. Run P (bar) T (ºC) C (g/ml) w1 w2 w3 1 70 30 0.05 0.9128 0.0823 0.0049 2 70 30 0.20 0.8996 0.0811 0.0193 3 90 30 0.05 0.9669 0.0312 0.0019 4 90 30 0.20 0.9616 0.0310 0.0074 5 70 40 0.05 0.8861 0.1075 0.0064 6 70 40 0.20 0.8694 0.1055 0.0251 7 90 40 0.05 0.9502 0.0470 0.0028 8 90 40 0.20 0.9423 0.0466 0.0111 In the case of using Limonene as solvent with the lowest concentration, thick, colourless and nonporous layers can be appreciated (Figure 7.18) combined with areas where foams apparently present higher density and nucleation bubbles appear to be not distributed homogeneously. The formation of a non-porous skin is due to the presence of gas molecules near the surface of the sample diffusing out of the sample faster than they can form nuclei. Rapid diffusion out of the sample creates a reduction of the layer near the surface in which the gas concentration is too low to contribute to nucleate and growth [51]. When Polystyrene initial concentration increased, the structures obtained are more rigid, compact and less rubbery that in the case of lower concentration. Similar trends were observed when cymene was used as solvent. Nevertheless when the initial concentration was 0.20 g PS/ml p-Cymene, the foam structure is more uniform and cells are more and smaller. This fact is mainly determined by the competition between bubble nucleation and growth rates [74]. At higher temperature, foams produced present larger cells because when the CO 2 concentration in the dissolution is low because the effect of pressure on the foam structure is negligible. In future works the study of the influence of the variables affecting the foaming process will be thoroughly studied and it will be checked the foam structures by means of scanning electronic microscopy (Figure 7.19). 220 Chapter 7 Figure 7.19. SEM images with 400x magnification of PS foam obtained at 90 bar, 40ºC and 0.20 g PS/ml Limonene. Finally, it is important to stand out that limonene and/or cymene are effectively removed from the PS and the foams obtained are completely free of solvent as was confirmed by TGA measurements. Figure 7.20 shows the weight loss profile of one of the foamed samples. The first of the weight losses at around 176 ºC, corresponds to the boiling point of the terpene oil. The second weight loss at 420 ºC corresponds to PS degradation. In this way, it was confirmed that the final solvent concentration in the foam is always lower than 5 % at the studied working conditions. 100 3.0 90 Weight (%) 70 2.5 2.0 60 50 1.5 40 1.0 30 20 0.5 Deriv. Weight (%/ºC) Weight Deriv. Weight 80 10 0 0 50 0.0 100 150 200 250 300 350 400 450 500 550 600 Temperature (ºC) Figure 7.20. Thermogravimetric analysis (TGA) and differential Termogravimetric Analysis (DTGA) of PS foam obtained at 90 bar, 30 ºC and 0.20 g/ml Cymene as initial concentration. 221 Supercritical Antisolvent and Foaming Processes 7.5. Statistical study of the foaming process of PS/p-Cymene solutions In this section, a full factorial two level design with 3 factors (pressure, temperature and concentration) and a center point (23+1) was proposed and nine runs were carried out to determine the influence of the mentioned variables on the foams cells size in order to optimize this parameter. The levels of each factor are indicated in Table 7.19. Table 7.19. Coded and uncoded matrix of the experiments. Contact time: 240 min, depressurization time: 1.5 min. Run p T C p (bar) T (ºC) C (g/ml) 1 -1 -1 -1 70 30 0.05 2 -1 -1 1 70 30 0.20 3 1 -1 -1 90 30 0.05 4 1 -1 1 90 30 0.20 5 0 0 0 80 35 0.125 6 -1 1 -1 70 40 0.05 7 -1 1 1 70 40 0.20 8 1 1 -1 90 40 0.05 9 1 1 1 90 40 0.20 The coding scheme used to describe the factor levels is based on the “1” and “−1” signs, where denote the high and low level of each factor and 0 corresponds to the mean between the two levels of all factors. The limits of the operating conditions were selected following the principles explained in the previous section: vanishing of the interfacial tension and temperature above glass transition temperature of the ternary mixture. As it was explained, among the most important solid-state properties defined by the foaming process are the dimensional properties of the final product (cells diameter and morphology). The variables studied in this work were the diameter of the pore cells, its standard deviation and the cells density to determine the foams structure and homogeneity. Results have been divided in three sections. Initially, the results of the foaming process are shown, next the characterisation of samples was performed to optimise the results during the last section. All experiments in the design matrix shown in Table 7.19 were carried out according to the same experimental procedure, with the operating conditions indicated by the coding scheme (high and low levels for each factor). Figure 7.21 shows the microphotographs analyzed by scanning electron microscopy where the structure and cell size of the foamed PS can be observed. The most 222 Chapter 7 homogeneous foams which presented higher cells density, smaller and well distributed pores were obtained at high level of pressure and concentration and low values of temperature. Figure 7.21. SEM images with 2000x magnification of PS foam obtained as a function of the initial concentration of PS, temperature and pressure. T:30ºC T:35ºC T:40ºC C0=0.05 g/ml C0=0.20 g/ml C0=0.125 g/ml C0=0.05 g/ml C0=0.20 g/ml p: 70 bar p: 85 bar No foam No foam No foam p: 90 bar From the microphotographs, data on the average diameter of cells data were fitted to a normal distribution curve. As it was shown n the previous section, the foaming process can not be performed at 70 bar and 40ºC nor at 85 bar and 35ºC, because it is not feasible and higher PS concentration solutions were obtained instead of foams [31]. The foaming process was not achieved under the mentioned conditions because the CO2 concentration was too low to solubilise fully p-Cymene. The solubility of CO2 in polymers increases under higher pressure and lower temperature [75] (see Chapter 5). Also, the CO2 density increases at high pressure and low temperature like its solvent power [76] so p-Cymene should be preferably removed at high densities. Furthermore, the critical points and the Tg of the mixtures was studied and results are shown in Figure 7.22. According to Figure 7.22, at 30 ºC foams were obtained independently of the working pressure since they are below the glass transition temperature of the mixture. Nevertheless, at higher temperature, the foaming process is achieved when the working pressure is above the critical pressure of the mixture, which demonstrates that p-Cymene is fully soluble in CO2 and the separation process can be performed. According to these reasons, runs 5, 6 and 7 provided the less favourable conditions to perform the foaming process. 223 Supercritical Antisolvent and Foaming Processes 110 100 Pres (bar) 90 80 70 60 Tg PS/p-Cymene solutions Critical mixing point of 0.05 g PS/ml p-Cymene solutions Critical mixing point of 0.20 g PS/ml p-Cymene solutions Foam No foam 50 40 25 30 35 40 45 Temp (ºC) Figure 7.22. Influence of glass transition temperature and phase equilibrium boundary on the feasibility of the foaming process: (▲) successful and/or () unsuccessful. (—) Decrease of Tg of the PS/p-Cymene solutions as a function of CO2 pressure; (…) critical point of the mixture at 0.05 g PS/ml p-Cymene and () 0.20 g PS/ml p-Cymene. 7.5.1. Characterisation All samples were characterized to determine the amount of residual solvent, their glass transition and the degradation temperature (Table 7.20). Table 7.20. Characterisation of the Polystyrene foams. Concentration of p-Cymene in the final products (% wt Cym), glass transition temperature (Tglass) and degradation temperature (Tdeg). Run p T C p (bar) T (ºC) C (g/ml) % wt Cym Tglass (ºC) Tdeg (ºC) 1 -1 -1 -1 70 30 0.05 2.70 104.29 415.54 2 -1 -1 1 70 30 0.2 5.00 104.04 416.85 3 1 -1 -1 90 30 0.05 4.00 101.81 415.82 4 1 -1 1 90 30 0.2 5.49 101.85 417.56 5 0 0 0 80 35 0.125 No foam 6 -1 1 -1 70 40 0.05 No foam 7 -1 1 1 70 40 0.2 No foam 8 1 1 -1 90 40 0.05 8.22 98.88 421.18 9 1 1 1 90 40 0.2 7.61 101.30 418.76 224 Chapter 7 The average concentration of p-Cymene present in PS foams is around 5% but the foams are not structurally affected because of the solvent traces. Lower temperatures enhance the solubility of p-Cymene in CO2 and consequently, pCymene traces in the final foam are reduced. Also, terpene oil traces would be easily removed from the foams by adding an extra stream of CO 2 once depressurization is performed. The degradation temperature and the glass transition temperature were kept almost constant in all samples compared with the original PS pellets (105.81 ºC and 408.92 ºC, respectively). An average deviation of 3.58% in the Tg and 2.13% in the Tdeg was obtained which confirmed that PS maintained its original calorimetric properties. In addition, the molecular weight distribution by means of GPC to confirm that fractionation of polymer do not occur during the experiments. Since the cell size of the final product is defined as one of the most important properties while the calorimetric and structural properties remain constant, the pore cells, its standard deviation and the cells density were studied (Table 7.21). Table 7.21. Average diameter of the pore cells (m), standard deviation and cells density (cells/cm3) of the PS foams. Run p T C p (bar) T (ºC) C (g/ml) Diameter Cells Density Std. Dev. (cells/cm3) m) 1 -1 -1 -1 70 30 0.05 4.472 2.264 2.40·109 2 -1 -1 1 70 30 0.20 10.899 4.549 5.52·108 3 1 -1 -1 90 30 0.05 7.228 2.560 2.04·1010 4 1 -1 1 90 30 0.20 2.822 1.275 2.47·1011 5 0 0 0 80 35 0.125 No foam 6 -1 1 -1 70 40 0.05 No foam 7 -1 1 1 70 40 0.20 No foam 8 1 1 -1 90 40 0.05 5.270 1.217 1.51·1010 9 1 1 1 90 40 0.20 20.345 5.838 6.45·108 Generally higher cells densities are achieved when the pore size are smaller, but in the cases in which the foams are heterogeneous, small cell diameters could lead to low values of cells density. According to Table 7.21, the smallest and narrowest distributed pores are obtained in run 4 which also shows the highest value of cells density. By contrast, the foams obtained in run 9 exhibit big pores, non homogeneously distributed and low values of cells density. 225 Supercritical Antisolvent and Foaming Processes 7.5.2. Optimisation Analysis of the experimental data is focused on the production of homogeneous foams with high cells density which means minimization of the cells size and its standard deviation and maximization of the cell density. With three samples not foaming it is difficult to draw any accurate conclusions using the traditional statistical analysis of variance to determine the input factors statistically significant and finding the optimum levels. Therefore, the influence of each variable on the cell structure is studied individually. The effect of pressure on the cell size and its distribution was studied at constant temperature (30 ºC) and concentration (0.20 g PS/ml p-Cymene). PS foams were produced at 70 and 90 bar (runs 2 and 4). The effect of pressure on the average pore diameter and its distribution is illustrated in Figure 7.23 where an increase in CO2 pressure tended to produce smaller and narrower distributed pores. 80 Runs Average Std. Dev. 2; P: 70 bar 10.90 4.549 4; P: 90 bar 2.822 1.275 70 Probability Density 60 T: 30 ºC; C0: 0.20 g PS/ml p-Cymene 50 40 30 20 10 0 0 5 10 15 Average Pore Cells 20 Figure 7.23. Effect of pressure on the cells size through runs 2 and 4 performed at 30 ºC, 0.20 g PS/ml p-Cymene, 70 and 90 bar, respectively. As the pressure increases, more CO2 was dissolved into the polymer solution and caused higher plasticization and viscosity reduction which facilitated the growth in the number of pores and reduced their size leading to higher rate of nucleation when pressure was reduced [8, 24, 25]. This fact was observed through the increase of cells density from 5.52·108 to 2.47·1011 cells/cm3. The influence of temperature on the porous structure was studied at constant pressure (90 bar) and concentration (0.20 g PS/ml p-Cymene) while temperature ranged between 30 and 40 ºC (runs 4 and 9) and is observed in Figure 7.24. 226 Chapter 7 140 Average Std. Dev. Runs 2.822 1.275 4; T: 30 ºC 20.34 5.838 9; T: 40 ºC 120 P: 90 bar; C0: 0.20 g PS/ml p-Cymene Probability Density 100 80 60 40 20 0 0 8 16 Average Pore Cells 24 32 Figure 7.24. Effect of temperature on the cells size through runs 4 and 9 performed at 90 bar, 0.20 g PS/ml p-Cymene, 30 and 40 ºC, respectively. It should be emphasized that the foaming process is hampered when the working temperature was above the glass transition temperature of the mixture and foams are only obtained at higher pressure. With higher temperatures (Figure 7.24), the cell size increases from 2 to 20 m in the studied range. The increase of the pore diameter is due to a reduction in the viscosity of the solution and a decrease of the CO2 solubility into the PS/p-Cymene solution which produces fewer nuclei but consequently larger cells [51, 58, 59, 77], i.e., a lower cell density. The effect of Polystyrene concentration on the cell size was studied depending on the pressure and temperature used because different trends were obtained (Figure 7.25). Runs 1-2 (70 bar and 30 ºC) and runs 8-9 (90 bar and 40 ºC) indicate an increase in the cell size when PS concentration increases from 4-5 m at 0.05 g PS/ml p-Cymene to 10-20 m, at 0.20 g PS/ml p-Cymene. An increase of the pore size could be mainly attributed to the lower CO2 solubility which is produced at low pressure and high temperature [75]. With higher initial concentration of the solution, the CO2 sorption decreases and the nucleation rate is lower as the amount of cells formed, but they are bigger and inhomogeneously distributed. Comparing runs 3 and 4, it can be stated that with higher PS concentration, the cell size decreases. At 90 bar and 30 ºC the solubility of CO 2 in the solution was improved and consequently nucleation is enhanced which causes the formation of many homogenously distributed nuclei. 227 Supercritical Antisolvent and Foaming Processes a) 50 Runs Average Std. Dev. 1; C: 0.05 g/ ml 4.472 2.264 2; C: 0.20 g/ ml 10.90 4.549 Probability Density 40 P: 70 bar; T: 30 ºC 30 20 10 0 0 5 10 15 20 Average Pore Cells (b) 140 Runs Average Std. Dev. 8; C: 0.05 g/ml 5.270 1.217 9; C: 0.20 g/ml 20.34 5.838 Probability Density 120 P: 90 bar; T: 40 ºC 100 80 60 40 20 0 4 8 12 16 20 24 28 32 Average Pore Cells Figure 7.25. (a) Effect of concentration on the cells size through runs 1 and 2 performed at 70 bar, 30 ºC, 0.05 and 0.20 g PS/ml p-Cymene, respectively. (b) Effect of concentration on the cells size through runs 8 and 9 performed at 90 bar, 40 ºC, 0.05 and 0.20 g PS/ml p-Cymene, respectively. In order to obtain some general conclusions about the effect of the working conditions over the cells size, the global effect of pressure, temperature and concentration was studied. The main purpose of the section is the foaming of PS/pCymene solutions in order to get small pores and homogeneous cells foams. According to Figures 7.23 - 7.25, the PS foams with the smallest and narrowest pore diameter together with the highest cell density are obtained at 90 bar, 30 ºC and 0.20 g PS/ml p-Cymene because a decrease of pressure produces an increase of the cell size while the pore diameter distribution remains practically constant. The optimum results are displayed in Figure 7.26 where the foams presented small and homogeneously distributed pores. 228 Chapter 7 (a) (b) Figure 7.26. Microphotographs of the optimum PS foams obtained at 90 bar, 30 ºC and 0.20 g PS/ml p-Cymene (a) 3000x magnification and (b) 6000x magnification To sum up, a proper foaming of Polystyrene under mild working conditions could be achieved and the structure of the foams produced can be tailored by altering the pressure and the temperature. But a rigorous study is shown in the subsequent section with the aim of study all the variables involved in the foaming process. 7.6. Rigorous Statistical study of the foaming process of Polystyrene/Limonene solutions As it has been explained in the previous section, during the foaming process, several parameters should be studied since they have influence on the final structure or characteristics of the foams. Among the most important solid-state properties defined by the foaming process are the dimensional properties of the final product (diameter and morphology). The responses studied in this work were: the diameter of the pore cells, its standard deviation (indicative of the foam homogeneity) and the cells density. In this research section, the pressure, temperature, initial concentration of the PS solution, contact time and depressurization time were considered as variable. Although the most frequently design of tests is a full factorial in order to minimize the number of tests required, fractional factorial experiments (FFEs) design of experiments was developed. In particular, a design called Taguchi method was applied by using an orthogonal array of L27, with only twenty seven experiments. The three levels for each factor are shown in Table 7.11, and the values were chosen according to previous studies on PS foaming from solutions [31]. 229 Supercritical Antisolvent and Foaming Processes Table 7.11. Factors studied in the design of experiments (DOE) of the foaming process and their levels. Pressure (bar) Temperature (ºC) Concentration (g/ml) Contact time (min) Depressurization time (min) Level1 Level2 Level3 70 30 0.10 60 1.5 80 35 0.15 150 15.45 90 40 0.20 240 30 Furthermore, statistical analysis of variance (ANOVA) was applied to determine the effect of the input factors that was statistically significant and to find the optimum levels. The statistical analysis performed with the experimental results was made using commercial software package Statgraphics Plus 5.1 (Manugistics, Inc. Rockville, MD, USA). The test of statistical significance, p-value, was determined accordingly to the total error criteria considering a confidence level of 95%. Results are divided in three sections. Initially, experimental data together with the results obtained from the design of experiments is analyzed. Next, a brief description of the Classical Nucleation Theory is applied and finally, the characterization of the foams is assessed. 7.6.1. Foaming process All experiments in the design matrix shown in Table 7.11 were carried out according to the same experimental procedure, with the operating conditions indicated by the coding and the runs were established under a randomness criterion. Table 7.12 shows the structure of Taguchi’s orthogonal array design and some representative microphotographs. The average cells diameter, their standard deviation and the cell density were calculated from the analysis of several microphotographs in each experiment. 230 70 70 70 70 70 70 70 70 70 2 3 4 5 6 7 8 9 40 40 40 35 35 35 30 30 30 0.20 0.20 0.20 0.15 0.15 0.15 0.10 0.10 0.10 240 150 60 240 150 60 240 150 60 30 30 30 15.45 15.45 15.45 1.5 1.5 1.5 p C tcontact tdep T (ºC) (bar) (g/ml) (min) (min) 1 Run Scanning Electron Microphotography No foams were obtained No foams were obtained No foams were obtained 42.473 3.531 Highly porous skin Heterogeneous distribution 4.18 Very well defined pore surface 2.932 4.259 Very well defined pore surface Noncontinuous foam, nonporous homogeneous skin Average (m) Observation 11.667 0.913 0.834 2.323 1.202 Std. Dev 3.134·106 2.540·109 4.599·109 3.930·109 7.051·109 Cells Density (cells/cm3) Table 7.12. L27 orthogonal array. Runs, experimental working conditions, microphotographs, diameter cells and their standard deviations and cells density. Chapter 7 231 232 p (bar) 80 80 80 80 80 80 Run 10 11 12 13 14 15 35 35 35 30 30 30 0.10 0.10 0.10 0.20 0.20 0.20 240 150 60 240 150 60 30 30 30 15.45 15.45 15.45 T C tcontact tdep (ºC) (g/ml) (min) (min) Scanning Electron Microphotography 8.82 8.051 3.821 30.057 43.71 7.5 Very homogeneous pore distribution Very homogeneous distribution and small size pores Nonporous skin, very few big pores Nonporous and rough skin, very few big pores Heterogeneous, rough and few pores skin 3.236 8.862 0.908 5.019 2.561 Average Std. Dev (m) Fine, brittle, medium size pores foam Observation 3.199·108 7.932·108 1.266·109 1.756·109 4.433·108 2.738·108 Cells Density (cells/cm3) Table 7.12 (continued) L27 orthogonal array. Runs, experimental working conditions, microphotographs, diameter cells and their standard deviations and cells density Supercritical Antisolvent and Foaming Processes p (bar) 80 80 80 90 90 90 Run 16 17 18 19 20 21 30 30 30 40 40 40 T (ºC) 0.15 0.15 0.15 0.15 0.15 0.15 240 150 60 240 150 60 30 30 30 1.5 1.5 1.5 C tcontact tdep (g/ml) (min) (min) Scanning Electron Microphotography Average (m) 57.116 5.747 25.695 3.939 9.918 6.982 Observation Non very porous skin, big size and heterogeneous pores Pores inside cavities, non homogeneous pores Low porosity, big size pores Very homogeneous and small pores Very well defined and big cells, high cells density Heteregeneous and non very porous skin 3.243 2.362 1.122 10.769 2.207 28.85 Std. Dev 7.363·1010 2.539·1010 4.867·1010 4.868·107 2.223·109 2.801·107 Cells Density (cells/cm3) Table 7.12 (continued) L27 orthogonal array. Runs, experimental working conditions, microphotographs, diameter cells and their standard deviations and cells density Chapter 7 233 234 p (bar) 90 90 90 90 90 90 Run 22 23 24 25 26 27 40 40 40 35 35 35 T (ºC) 0.10 0.10 0.10 0.20 0.20 0.20 240 150 60 240 150 60 15.45 15.45 15.45 1.5 1.5 1.5 C tcontact tdep (g/ml) (min) (min) Scanning Electron Microphotography Average (m) 8.82 8.051 3.821 30.057 43.71 7.5 Observation Fine, brittle, medium size pores foam Very homogeneous pore distribution Very homogeneous distribution and small size pores Nonporous skin, very few big pores Nonporous and rough skin, very few big pores Heterogeneous, rough and few pores skin 3.236 8.862 0.908 5.019 2.561 Std. Dev 3.199·108 7.932·108 1.266·109 1.756·109 4.433·108 2.738·108 Cells Density (cells/cm3) Table 7.12 (continued) L27 orthogonal array. Runs, experimental working conditions, microphotographs, diameter cells and their standard deviations and cells density Supercritical Antisolvent and Foaming Processes Chapter 7 According to the microphotographs shown in Table 7.12 the most homogeneous foams which presented higher cells density, smaller size and are well distributed are obtained at low temperature and high concentration. From the analysis of microphotographs, the mean diameter of cells data were fitted to a normal distribution curve and the selected runs are shown in Figure 7.27. The smallest pores with the narrowest distribution were obtained from runs 3, 5 and 19. 0.5 Run 3 13 19 21 Probability Density 0.4 Average Std. Dev. 3.531 0.8344 30.06 8.862 3.939 1.122 6.982 3.243 0.3 0.2 0.1 0.0 0 8 16 24 Mean Cell Size 32 40 48 Figure 7.27. Probability density function of average cell pore diameter (m) for the runs ( ) 3, ( ) 13, ( ) 19 and ( ) 21. It is important to highlight that at 70 bar and 40ºC, no foams were obtained as it was previously observed [31]. To try to explain the lack of foams during the runs 79, the evolution of the working conditions as well as the Tg modification during the process were studied (Figure 7.28). The left axis represents the variation of the temperature (operating and glass transition temperature) along the experiments while the right axis shows the evolution of pressure. In the cited runs, the Tg is lower than the working temperature because the plasticising effect of CO 2, so the crystallization of polymer chains is practically not expected, unless the state of the solution is supercooled and the rate of crystallization high enough. The rest of the experiments presented values of Tg higher than different the operating temperature, which allows the nucleation of bubbles and foaming of the polymer. 235 Supercritical Antisolvent and Foaming Processes 100 Temp (ºC) 75 80 60 50 40 25 20 0 0 0 10 20 30 40 50 60 70 80 90 100 100 80 80 60 60 40 40 20 20 0 0 0 25 50 75 100 125 150 175 100 100 Run 9 60 60 40 40 20 20 0 0 25 50 75 100 125 150 175 200 225 250 Pressure (bar) 80 80 0 Pressure (bar) Temp (ºC) 100 Run 8 Temp (ºC) Pressure (bar) 100 Run 7 275 Time (min) Figure 7.28. Evolution of the operating conditions during the foaming process during the runs 7, 8 and 9. Evolution of temperature (dashed lines, left axis): glass transition temperature (■), and working temperature (○) and pressure (solid line, right axis). According to Figure 7.28, it could be concluded that a gap between the working temperature and the Tg of PS is required to promote the polymer crystallization and foaming. Taguchi approach involves two steps: the proper selection of the orthogonal arrays and its analysis of variance (ANOVA). Thus, ANOVA determines the contribution of each parameter on the responses factors (Table 7.13). The influence factor will be significant if the value of critical level (p) is lower than 0.05. As can be observed from p-value, the effects statistically significant over the mean pore diameter are the temperature, the concentration and the pressure while the contact time and the depressurization time are not statistically significant with a p-value higher than 0.05. The standard distribution of the pore cells was also affected by the previously mentioned variables. Nevertheless, the pressure is the only effect which affects significantly the cells density. 236 Chapter 7 Table 7.13. Main effects and interactions from L27 orthogonal array on the mean pore cell diameter, the standard deviation and the cells density. Factor Cells Diameter (m) Std. Dev. Cells Density (cells/cm3) p effect (±error std.) -27.55 (±10.92) -30.40 (±12.69) 7.53·1010 (±3.54·1010) p-value 0.0197 0.0281 0.0456 T effect (±error std.) 56.80 (±10.92) 50.89 (±12.69) -1.75·1010 (±3.54·1010) p-value 0.0001 0.0007 0.6258 C effect (±error std.) 32.12 (±10.92) 33.66(±12.69) -4.56·1010 (±3.54·1010) p-value 0.0078 0.0182 0.2120 Contact time effect (±error std.) 7.09 (±10.92) 13.99(±12.69) 5.80·1010 (±3.54·1010) p-value 0.5233 0.2979 0.1168 Dep time effect (±error std.) -12.28 (±10.92) -12.83(±12.69) 4.09·1010 (±3.54·1010) p-value 0.2731 0.3238 0.2609 According to the results shown in Table 7.13, the morphology of the foams is mainly affected by the pressure and the temperature. The standardized effects of the independent variables on the cells diameter, its standard distribution and the cells density were investigated by mean of a Pareto chart (Figure 7.29). The length of each bar indicates the standardized effect of the selected factor on the different responses and its colour represents if the contribution was positive or negative. The positive effects (grey colour) presented a favourable effect on the response while the negative effects (black bars) shown an antagonistic effect on it. 237 Supercritical Antisolvent and Foaming Processes (a) (b) (c) Figure 7.29. Pareto chart showing the standardized effect of independent variables (A: pressure; B: temperature; C: concentration; D: contact time; E: venting time) on the average cells diameter (a), its standard deviation (b) and cells density (c). According to Figure 7.29 (a) in the case of the average cell pore diameter, the temperature and depressurization time promote the increase of the pore sizes, whereas the pressure favours the formation of smaller pore cells. In Figure 7.29 (b) the significantly influence of pressure, temperature, concentration and depressurization time on the standard distribution of the pore cells was shown. Only an increase of pressure facilitates the homogeneity of the foam while an increase of the mentioned variables tends to increase the standard deviation of the pore cells diameter. Finally, Figure 7.29 (c) shows that only the pressure significantly affects the cells density. This fact supports the non homogeneity of the foams, since the cell density should be significantly influenced by the same variables than the cell pore diameter, but they should indicate opposite effects on the studied response. Next, the effect of each factor in the foaming characteristics is discussed. 238 Chapter 7 The effect of pressure on the average diameter of the cells was studied at 70, 80 and 90 bar. With increasing pressure, the mean pore diameter decreases and it can be found that if the cell size decreases, the amount of cells increases with pressure. They became narrower distributed with increasing pressure, which means that more uniform foam structure could be obtained at higher pressure [51, 58-60, 77]. At higher pressure, more CO2 is dissolved into the polymer solution and caused higher plasticization and viscosity reduction (see Chapters 5 and 6) [78]. These facts imply higher amount of CO2 used to nucleation which facilitates the growth of nucleis and results in more and smaller cells (higher cells density and lower pores diameter). The influence of the working temperature was investigated between 30 and 40ºC. An increase in the temperature of the sample produces an increase in the mean pore diameter as well as an increase in its heterogeneity and a decrease in the cells density [51, 77]. With increasing temperature, the CO2 solubility in the polymer solution decreases. Since at higher temperatures there is less dissolved fluid into the polymer matrix available for the nucleation and growth of pores, fewer nuclei (that should share this fluid) are formed [59, 77]. Higher temperature lower the viscosity of the matrix and increase the diffusion rate of CO 2, rendering the cell growth faster, thus a narrow cell size distribution could be achieved at lower processing temperature. The effect of concentration on the average diameter of the cells was examined in the range of 0.1 to 0.2 gPS/ml Limonene. According to Table 7.13 and Figure 7.29, increasing the initial concentration of PS in the solution, the mean pore diameter and the heterogeneity increase. Concentration is not statistically significant over the cells density, nevertheless a decrease in the cells density was observed when concentration of Polystyrene increases. With increasing the initial concentration of the solution, the CO2 sorption decreases as it was observed experimentally in Figure 5.7, so the nucleation rate is lower as the amount of cells formed, but they are bigger and non homogeneously distributed. The contact time of CO2 in PS/Limonene solutions was studied in a range between 60 and 240 min according previous literature references [51] An increase in the contact time produces more uniform structures and smaller pore cells. Although Reverchon’s studies concluded that 150 min was enough to promote homogeneous microcellular structure, we observe that when Limonene is presented in the mixture, the minimum contact time increases up to 240 minutes. Finally, the effect of depressurization time was considered in the study of PS/Limonene solution foaming. Fast (1.5 min) or slow (30 min) depressurization time was tuned to determine its influence on the mean pore cells. The mean pore diameter decreases when the depressurization time increased. At high depressurization time (low depressurization rates) the CO 2 entrapped in the swelling solution is slowly removed, decreasing the cells size and increasing the cell population. Nevertheless, high depressurization rate produces the coalescence of neighbour cells as a consequence of the abrupt elimination of CO 2 [59, 79]. 239 Supercritical Antisolvent and Foaming Processes In order to obtain some general conclusions about the effects over the studied responses, Figure 7.30 summarized the response surfaces of the average mean pore of the cells and their standard deviation. In this case the cells density is not considered because only one factor (the pressure) was statistically significant. Superficie de Respuesta Estimada Concentration=0.15,Contact time=150.0,Dep time=15.75 Mean Cells Diameter (a) 100 80 60 40 20 40 0 70 Mean Cells Diameter 0.0 12.0 24.0 36.0 48.0 60.0 72.0 84.0 96.0 108.0 120.0 35 80 Pressure (bar) 90 30 Temperature (ºC) Superficie de Respuesta Estimada Concentration=0.15,Contact time=150.0,Dep time=15.75 Std. Dev. (b) 90 75 60 45 30 15 0 40 Standard Deviation -20.0 -8.0 4.0 16.0 28.0 40.0 52.0 64.0 76.0 88.0 100.0 112.0 35 30 70 80 Temperature (ºC) 90 Pressure (bar) Figure 7.30. Response surfaces showing the relation between the pore cells (a) and its standard deviation (b) as a function of pressure and temperature. According to Figure 7.30 (a), an increase in pressure and a decrease in temperature at constant concentration, contact time and depressurization time acted increasing the sorption of the supercritical solvent into the polymeric matrix which decreases the pore size while increases its homogeneity. Whereas, as it is observed in Figure 7.30 (b) the heterogeneity (higher standard deviation) of the pores sizes was obtained at higher temperatures and lower pressures. 7.6.2. Nucleation mechanisms The main objective of any foaming process is to obtain foams with high cells density and low cell size which implies high nucleation rate. Nevertheless, the formation of nuclei in a viscous liquid is a complex mechanism influenced by several parameters 240 Chapter 7 [70]. During this section the main factors which affected the formation of microcellular foams are determined. Various mechanisms of nucleation have been proposed in the literature for nucleation of gas bubbles in thermoplastics, including homogeneous, heterogeneous, mixed-mode, shear-induced and void nucleation theories. However, none of these theories give quantitative predictions for even the simplest cases of bubble nucleation in polymeric liquids. One of the key processes to control the cellular structure is the nucleation and growth of gas cells dispersed throughout the polymer. The resulting structure depends on a balance between several mechanisms, such as diffusion of CO2 in the polymer, the solubility and the plasticization effect, which depends intrinsically of the polymer and the saturation conditions [80]. The nucleation theory can be used to describe the thermodynamic and kinetic effects on the foam generation. The Classical Nucleation Theory (CNT) has been selected to explain some of the experimental facts obtained because it is still regarded as the most successful approach for prediction of nucleation phenomena [81]. The steady state nucleation rate, defined as the number of critical nuclei formed per unit time in a unit volume of the parent phase is given by: [7.13] where Jss is the number of critical nuclei formed per unit time in a unit volume, W is the reversible work of critical nucleus formation, k B is the Boltzmann factor, and T is the absolute temperature. J0, according to the kinetic consideration of the bubble formation process, is defined as: [7.14] where N is the number of gas molecules per unit volume, it means the density of the blowing agent molecule in polymer-blowing agent solutions, m is the mass of the gas molecule, and is the surface tension between the metastable parent phase and the nucleating phase. B is a dimensionless factor that varies between 2/3 and 1, depending on the viscosity of the material; in the case of polymer is considered to have a value of 1. W presents a larger impact on the value of the nucleation rate and its accurate calculation is considered more essential that the proper calculation of J0. In the cases where p is large, W becomes close to zero and the exponential term of the equation which represents the steady state nucleation rate (Eq. [7.13]) is mostly affected by the pre-exponential factor J0. For a homogeneous nucleation, W is evaluated to be: [7.15] 241 Supercritical Antisolvent and Foaming Processes where P is the difference between Pnucleus, the pressure of the metastable phase (inside the gas nucleus) and Psystem, the pressure of the nucleating phase (polymer/gas solution) [56]. P is the real driven force of the foaming process and due to the importance of its accurate determination it is calculated from the experimental data obtained because it is difficult to obtain experimentally. Next, the calculated drop pressure could be used to predict the behaviour of the foam using the Classical Nucleation Theory. According to equations 7.13, 7.14 and 7.15 and considering the experimental data shown, the value of P required to obtain the critical nuclei (JSS) is calculated. N can be obtained from the equilibrium data shown in Chapter 5, was determined experimentally and it was shown in our previous work [31] and in Chapter 6 and B was fixed at 2/3 because the solutions were not very viscous. In those cases where foams are not obtained, an arbitrary value of 109 cells/cm3 was selected as cell densities because of the microcellular foams definition. Results are shown in Table 7.14. According to Table 7.14 P decreases when pressure increases and temperature decreases. The main variables that present influence over the P are the interfacial tension between phases and the temperature. The interfacial tension of Polystyrene in Limonene solution decreases when the pressure of CO2 increases, by this reason, P required to foam also decreases. The small range of concentration does not influence significantly on the interfacial tension and consequently it is not affected by far the minimum pressure required to foam. The contact time promoted the CO 2 sorption into the polymeric solution. Nevertheless, this parameter does not affect the interfacial tension, because it is checked that a rapid achievement of the equilibrium composition is got between the drop and the interphase [31]. In general, minimum values of P are achieved when interfacial tension is close to the vanishing point. In these cases, the concentration of CO 2 in the polymer-solvent phase increase promoting a decrease of interfacial tension since the two phases in contact become more similar. The most suitable conditions to carry out the foaming process are those in which P is lower, because a very low driven force is necessary to ease nucleation and cell growth. In Table 7.14, the lowest values of P are observed at runs 10-12, where low and homogeneous cell sizes are obtained. 242 Chapter 7 Table 7.14. P required to carry out the foaming process. Run p (bar) T (ºC) C (g/ml) tcontact (min) tdep (min) P (bar) 1 70 30 0.10 60 1.5 17.23 2 70 30 0.10 150 1.5 17.05 3 70 30 0.10 240 1.5 17.10 4 70 35 0.15 60 15.45 22.23 5 70 35 0.15 150 15.45 22.59 6 70 35 0.15 240 15.45 20.32 7 70 40 0.20 60 30 35.46 8 70 40 0.20 150 30 35.46 9 70 40 0.20 240 30 35.46 10 80 30 0.20 60 15.45 0.00381 11 80 30 0.20 150 15.45 0.00385 12 80 30 0.20 240 15.45 0.00395 13 80 35 0.10 60 30 2.72 14 80 35 0.10 150 30 2.69 15 80 35 0.10 240 30 2.65 16 80 40 0.15 60 1.5 6.15 17 80 40 0.15 150 1.5 6.61 18 80 40 0.15 240 1.5 6.20 19 90 30 0.15 60 30 0.00426 20 90 30 0.15 150 30 0.00420 21 90 30 0.15 240 30 0.00430 22 90 35 0.20 60 1.5 0.00426 23 90 35 0.20 150 1.5 0.00420 24 90 35 0.20 240 1.5 0.00444 25 90 40 0.10 60 15.45 0.00397 26 90 40 0.10 150 15.45 0.00385 27 90 40 0.10 240 15.45 0.00376 The difference between the initial interfacial tension of the solutions and the interfacial tension of the polymer (without Limonene, only CO 2) at the same working conditions are represented in Figure 7.31. As it was previously observed, the lowest values of interfacial tension promote the polymer solutions foaming. However, the highest values of interfacial tension of solutions give the lowest gap 243 Supercritical Antisolvent and Foaming Processes between the initial interfacial tension of solutions and the interfacial tension of the polymer at the same pressure. The impossibility to process the foams from the solution at those conditions (P: 70 bar, T:40ºC) could also be attributed to the high driven force required. 0.0400 Initial Surface Tension (Solution) Final Surface Tension (Polystyrene) 0.0375 0.0350 (N/m) 0.0325 0.0300 0.0050 0.0025 0.0000 0 3 6 9 12 15 18 21 24 27 Run Figure 7.31. Interfacial tension of Polystyrene solutions (-■-) and Polystyrene (-○-) in the different runs. On the other hand, the influence of temperature on P was given by the inverse of the square root, thus, when temperature increased, P should decrease linearly which would help the foaming of the solutions at higher temperature. Nevertheless, experimental data do not show the expected trend. It is important to consider that when temperature increases CO2 sorption in PS or its solution decreases and the minimum driven force required to foam is increased because lower amounts of the gas are imbedded into the polymeric matrix. This fact confirms that lower temperature values are desired to foam the Polystyrene solutions. 7.6.3. Characterisation All samples (except those on which foaming were not produced) were characterized to determine the amount of residual solvent, their glass transition (Tg) and degradation temperature (Tdeg). Virgin Polystyrene pellets were also characterized in order to compare the results obtained from the analysis of the processed polymeric foams (Table 7.15). 244 Chapter 7 Table 7.15. Characterization of the Polystyrene foams. Concentration of Limonene in the final products (% wt Lim), glass transition temperature (Tglass) and degradation temperature (Tdeg). Run p (bar) T (ºC) C (g/ml) % wt Lim Tglass (ºC) Tdeg (ºC) 1 70 30 0.10 2.43 104.43 414.35 2 70 30 0.10 2.97 104.15 416.72 3 70 30 0.10 7.03 103.92 416.97 4 70 35 0.15 5 70 35 0.15 7.61 96.25 420.06 6 70 35 0.15 6.95 101.52 420.57 7 70 40 0.20 No observable foaming 8 70 40 0.20 No observable foaming 9 70 40 0.20 No observable foaming 10 80 30 0.20 2.78 104.08 415.81 11 80 30 0.20 2.12 102.28 414.77 12 80 30 0.20 3.54 95.15 414.55 13 80 35 0.10 6.58 103.06 421.97 14 80 35 0.10 6.38 94.57 415.97 15 80 35 0.10 6.74 104.46 416.90 16 80 40 0.15 6.34 101.65 421.14 17 80 40 0.15 4.37 100.78 416.46 18 80 40 0.15 7.84 95.91 420.07 19 90 30 0.15 3.56 101.00 415.83 20 90 30 0.15 4.44 102.62 415.81 21 90 30 0.15 6.53 101.07 419.30 22 90 35 0.20 4.59 23 90 35 0.20 5.36 24 90 35 0.20 2.66 104.14 417.33 25 90 40 0.10 8.52 76.62 423.40 26 90 40 0.10 7.91 101.13 418.95 27 90 40 0.10 7.30 101.46 418.57 105.81 408.92 Virgin Polystyrene pellets No observable foaming The amount of Limonene embedded in the foams os studied and results are shown in Table 7.15. The average concentration of solvent present in PS foams is around 245 Supercritical Antisolvent and Foaming Processes 5%, although there are some runs where Limonene amount is higher, the foams are not structurally affected because of the traces of solvent. Limonene would be easily removed from the foams adding an extra stream of CO 2 once depressurization is carried out because it has been assessed that Limonene is fully miscible in CO 2 at the working conditions (Chapter 6) [25]. According to Table 7.15, the degradation temperature and the glass transition temperature are kept almost constant in all samples. Low values of standard deviation confirm that polymer degradation or structural modification do not occur during the foaming process with CO 2 at high pressure. The decrease of Tg values in some cases can be attributed to the presence of Limonene which acts as plasticizer and consequently promotes the mobility of the polymeric chains decreasing its glass transition temperature. Finally, a comparison between commercial Polystyrene wastes (particularly, XPS which is used as insulation material) and the foams produced in our research group was accomplished (Figure 7.32). Average Pore (m) Cells Density (cells/cm3) Density (kg/m3) Recycled Polystyrene Commercial Polystyrene 3-20 1·109-1011 304.02 18.34 400-600 3-5 ·104 35 10 Figure 7.32. Comparison between the recycled Polystyrene (left) and the original wastes (right). As it is observed, in the recycled Polystyrene obtained in this research work, the cells are smaller, more homogeneous and consequently it presents higher cells density as well as a significant increase on bulk density. According to the values obtained it can be assessed that the production of microcellular foams from Polystyrene solutions is a feasible process. On that way, a very high added value product is obtained because of the improvement of the structural properties which is responsible of higher impact strength, toughness and better thermal stability. The foaming of Polystyrene at mild working conditions can be achieved and the structure of the foams produced can be tailored by altering mainly the pressure and the temperature. The use of a terpene solvent to perform an initial solution provided 246 Chapter 7 an increase in CO2 sorption which was the responsible of the plasticization of the Polystyrene. The polymer properties modification (decrease of glass transition temperature, interfacial tension, viscosity…), as a consequence of gas sorption, increases the CO2 diffusion coefficient which also benefited the nucleation and cell growth. According to the design of experiments analysis, the most suitable conditions to carry out the process of foaming Polystyrene/Limonene solutions were higher pressures (90 bar), low temperature (30ºC), low concentration (0.1 gPS/ml Limonene), high contact time (240 min) and depressurization time (30 min). At these conditions, the cell size decreased while the cell density increases which provided foams with high quality properties. The characterization of the foams showed that all the properties of the initial polymer were kept constant. 247 Supercritical Antisolvent and Foaming Processes References [1] E. Carretier, E. Badens, P. Guichardon, O. Boutin, G. Charbit, Hydrodynamics of supercritical antisolvent precipitation: Characterization and influence on particle morphology, Industrial and Engineering Chemistry Research, 42 (2003) 331-338. [2] C. Chen, S. Zhan, M. Zhang, Z. Liu, Z. Li, Preparation of poly(L-lactide) microparticles by a supercritical antisolvent process with a mixed solvent, Journal of Applied Polymer Science, 124 (2012) 3744-3750. [3] Y. Park, C.W. Curtis, C.B. Roberts, Formation of nylon particles and fibers using precipitation with a compressed antisolvent, Industrial and Engineering Chemistry Research, 41 (2002) 1504-1510. [4] Y. Pérez De Diego, F.E. Wubbolts, G.J. Witkamp, T.W. De Loos, P.J. Jansens, Measurements of the phase behaviour of the system dextran/DMSO/CO 2 at high pressures, Journal of Supercritical Fluids, 35 (2005) 1-9. [5] K. Wu, J. Li, Precipitation of a biodegradable polymer using compressed carbon dioxide as antisolvent, Journal of Supercritical Fluids, 46 (2008) 211-216. [6] J.O. Werling, P.G. Debenedetti, Numerical modeling of mass transfer in the supercritical antisolvent process: Miscible conditions, Journal of Supercritical Fluids, 18 (2000) 11-24. [7] J.O. Werling, P.G. Debenedetti, Numerical modeling of mass transfer in the supercritical antisolvent process, Journal of Supercritical Fluids, 16 (1999) 167-181. [8] D.J. Dixon, G. Luna-Bárcenas, K.P. Johnston, Microcellular microspheres and microballoons by precipitation with a vapour-liquid compressed fluid antisolvent, Polymer, 35 (1994) 3998-4005. [9] E. Reverchon, Supercritical antisolvent precipitation of micro- and nanoparticles, Journal of Supercritical Fluids, 15 (1999) 1-21. [10] C. Vemavarapu, M.J. Mollan, M. Lodaya, T.E. Needham, Design and process aspects of laboratory scale SCF particle formation systems, International Journal of Pharmaceutics, 292 (2005) 1-16. [11] E. Franceschi, M.H. Kunita, M.V. Tres, A.F. Rubira, E.C. Muniz, M.L. Corazza, C. Dariva, S.R.S. Ferreira, J.V. Oliveira, Phase behavior and process parameters effects on the characteristics of precipitated theophylline using carbon dioxide as antisolvent, Journal of Supercritical Fluids, 44 (2008) 8-20. [12] E. Reverchon, I. De Marco, Supercritical antisolvent precipitation of Cephalosporins, Powder Technology, 164 (2006) 139-146. [13] A. Martín, M.J. Cocero, Numerical modeling of jet hydrodynamics, mass transfer, and crystallization kinetics in the supercritical antisolvent (SAS) process, Journal of Supercritical Fluids, 32 (2004) 203-219. [14] S.D. Yeo, E. Kiran, Formation of polymer particles with supercritical fluids: A review, Journal of Supercritical Fluids, 34 (2005) 287-308. [15] Y.W. Kho, D.S. Kalika, B.L. Knutson, Precipitation of Nylon 6 membranes using compressed carbon dioxide, Polymer, 42 (2001) 6119-6127. [16] Y.P. De Diego, H.C. Pellikaan, F.E. Wubbolts, G.J. Witkamp, P.J. Jansens, Operating regimes and mechanism of particle formation during the precipitation of polymers using the PCA process, Journal of Supercritical Fluids, 35 (2005) 147-156. [17] A. Tenorio, P. Jaeger, M.D. Gordillo, C.M. Pereyra, E.J.M. De La Ossa, On the selection of limiting hydrodynamic conditions for the Supercritical AntiSolvent (SAS) process, Industrial and Engineering Chemistry Research, 48 (2009) 92249232. 248 Chapter 7 [18] P. Subra, C.G. Laudani, A. Vega-González, E. Reverchon, Precipitation and phase behavior of theophylline in solvent-supercritical CO2 mixtures, Journal of Supercritical Fluids, 35 (2005) 95-105. [19] R. Adami, L.S. Osséo, R. Huopalahti, E. Reverchon, Supercritical AntiSolvent micronization of PVA by semi-continuous and batch processing, Journal of Supercritical Fluids, 42 (2007) 288-298. [20] C.M. Hansen, Hansen solubility parameters: a user's handbook, CRC Press, 2000. [21] Y. Marcus, Are solubility parameters relevant to supercritical fluids?, Journal of Supercritical Fluids, 38 (2006) 7-12. [22] I. Gracia, J.F. Rodríguez, A. De Lucas, M.P. Fernandez-Ronco, M.T. García, Optimization of supercritical CO2 process for the concentration of tocopherol, carotenoids and chlorophylls from residual olive husk, Journal of Supercritical Fluids, 59 (2011) 72-77. [23] M.P. Fernández-Ronco, C. Ortega-Noblejas, I. Gracia, A. De Lucas, M.T. García, J.F. Rodríguez, Supercritical fluid fractionation of liquid oleoresin capsicum: Statistical analysis and solubility parameters, Journal of Supercritical Fluids, 54 (2010) 22-29. [24] J. Rincón, P. Cañizares, M.T. García, I. Gracia, Regeneration of used lubricant oil by propane extraction, Industrial and Engineering Chemistry Research, 42 (2003) 4867-4873. [25] H. Sovová, R.P. Stateva, A.A. Galushko, Essential oils from seeds: Solubility of limonene in supercritical CO2 and how it is affected by fatty oil, Journal of Supercritical Fluids, 20 (2001) 113-129. [26] C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. de Lucas, M.T. García, High-pressure phase equilibria of Polystyrene dissolutions in Limonene in presence of CO 2, The Journal of Supercritical Fluids, 84 (2013) 211-220. [27] A.T. Griffith, Y. Park, C.B. Roberts, Separation and recovery of nylon from carpet waste using a supercritical fluid antisolvent technique, Polymer-Plastics Technology and Engineering, 38 (1999) 411-431. [28] M. Rantakylä, M. Jäntti, O. Aaltonen, M. Hurme, The effect of initial drop size on particle size in the supercritical antisolvent precipitation (SAS) technique, Journal of Supercritical Fluids, 24 (2002) 251-263. [29] D.L. Obrzut, P.W. Bell, C.B. Roberts, S.R. Duke, Effect of process conditions on the spray characteristics of a PLA + methylene chloride solution in the supercritical antisolvent precipitation process, Journal of Supercritical Fluids, 42 (2007) 299-309. [30] A. Martín, M.J. Cocero, Micronization processes with supercritical fluids: Fundamentals and mechanisms, Advanced Drug Delivery Reviews, 60 (2008) 339350. [31] C. Gutiérrez, J.F. Rodríguez, I. Gracia, A. De Lucas, M.T. García, Development of a strategy for the foaming of polystyrene dissolutions in scCO 2, Journal of Supercritical Fluids, 76 (2013) 126-134. [32] B.E. Poling, J.M. Prausnitz, J.P. O'Connell, The properties of gases and liquids, McGraw-Hill, New York, 2001. [33] K. Lucas, Die Druckabhängigkeit der Viskosität von Flüssigkeiten – eine einfache Abschätzung, Chemie Ingenieur Technik, 53 (1981) 959-960. [34] E. Bender, Equations of state for ethylene and propylene, Cryogenics, 15 (1975) 667-673. 249 Supercritical Antisolvent and Foaming Processes [35] E. Reverchon, I. De Marco, E. Torino, Nanoparticles production by supercritical antisolvent precipitation: A general interpretation, Journal of Supercritical Fluids, 43 (2007) 126-138. [36] K.B. Deshpande, W.B. Zimmerman, Simulation of interfacial mass transfer by droplet dynamics using the level set method, Chemical Engineering Science, 61 (2006) 6486-6498. [37] A.D. Randolph, M.A. Larson, Theory of particulate processes: analysis and techniques of continuous crystallization, Academic Press, 1988. [38] E. Reverchon, R. Adami, G. Caputo, I. De Marco, Spherical microparticles production by supercritical antisolvent precipitation: Interpretation of results, Journal of Supercritical Fluids, 47 (2008) 70-84. [39] H. Chao-Hong, Y. Yong-Sheng, Estimation and infinite-dilution diffusion coefficients in supercritical fluids, Industrial and Engineering Chemistry Research, 36 (1997) 4430-4433. [40] J.L. Bueno, J.J. Suárez, I. Medina, Experimental binary diffusion coefficients of benzene and derivatives is supercritical carbon dioxide and their comparison with the values from the classic correlations, Chemical Engineering Science, 56 (2001) 4309-4319. [41] C. Mantell, M. Rodríguez, E. Martínez De La Ossa, Estimation of the diffusion coefficient of a model food dye (malvidin 3,5-diglucoside) in a high pressure CO2 + methanol system, Journal of Supercritical Fluids, 29 (2004) 165-173. [42] T. Fadli, A. Erriguible, S. Laugier, P. Subra-Paternault, Simulation of heat and mass transfer of CO2-solvent mixtures in miscible conditions: Isothermal and nonisothermal mixing, Journal of Supercritical Fluids, 52 (2010) 193-202. [43] C.H. He, Y.S. Yu, New equation for infinite-dilution diffusion coefficients in supercritical and high-temperature liquid solvents, Industrial and Engineering Chemistry Research, 37 (1998) 3793-3798. [44] F. Rindfleisch, T.P. DiNoia, M.A. McHugh, Solubility of polymers and copolymers in supercritical CO2, Journal of Physical Chemistry, 100 (1996) 1558115587. [45] M.L. O'Neill, Q. Cao, M. Fang, K.P. Johnston, S.P. Wilkinson, C.D. Smith, J.L. Kerschner, S.H. Jureller, Solubility of homopolymers and copolymers in carbon dioxide, Industrial and Engineering Chemistry Research, 37 (1998) 3067-3079. [46] J. Crank, The Mathematics of Diffusion, Clarendon Press, 1979. [47] M. Lora, A. Bertucco, I. Kikic, Simulation of the semicontinuous supercritical antisolvent recrystallization process, Industrial and Engineering Chemistry Research, 39 (2000) 1487-1496. [48] M. Mukhopadhyay, S.V. Dalvi, Mass and heat transfer analysis of SAS: Effects of thermodynamic states and flow rates on droplet size, Journal of Supercritical Fluids, 30 (2004) 333-348. [49] J.B. Bao, T. Liu, L. Zhao, G.H. Hu, X. Miao, X. Li, Oriented foaming of polystyrene with supercritical carbon dioxide for toughening, Polymer (United Kingdom), (2012). [50] D.L. Tomasko, A. Burley, L. Feng, S.K. Yeh, K. Miyazono, S. Nirmal-Kumar, I. Kusaka, K. Koelling, Development of CO2 for polymer foam applications, Journal of Supercritical Fluids, 47 (2009) 493-499. [51] E. Reverchon, S. Cardea, Production of controlled polymeric foams by supercritical CO2, Journal of Supercritical Fluids, 40 (2007) 144-152. 250 Chapter 7 [52] D. Sanli, S.E. Bozbag, C. Erkey, Synthesis of nanostructured materials using supercritical CO2: Part I. Physical transformations, Journal of Materials Science, 47 (2012) 2995-3025. [53] H. Zhong, Z. Xi, T. Liu, Z. Xu, L. Zhao, Integrated process of supercritical CO 2assisted melt polycondensation modification and foaming of poly(ethylene terephthalate), Journal of Supercritical Fluids, 74 (2013) 70-79. [54] L. Sorrentino, E. Di Maio, S. Iannace, Poly(ethylene terephthalate) foams: Correlation between the polymer properties and the foaming process, Journal of Applied Polymer Science, 116 (2010) 27-35. [55] D.C. Li, T. Liu, L. Zhao, W.K. Yuan, Foaming of linear isotactic polypropylene based on its non-isothermal crystallization behaviors under compressed CO 2, Journal of Supercritical Fluids, 60 (2011) 89-97. [56] Z. Guo, A.C. Burley, K.W. Koelling, I. Kusaka, L.J. Lee, D.L. Tomasko, CO2 bubble nucleation in polystyrene: Experimental and modeling studies, Journal of Applied Polymer Science, 125 (2012) 2170-2186. [57] J.-B. Bao, T. Liu, L. Zhao, G.-H. Hu, A two-step depressurization batch process for the formation of bi-modal cell structure polystyrene foams using scCO2, The Journal of Supercritical Fluids, 55 (2011) 1104-1114. [58] R. Liao, W. Yu, C. Zhou, Rheological control in foaming polymeric materials: I. Amorphous polymers, Polymer, 51 (2010) 568-580. [59] I. Tsivintzelis, A.G. Angelopoulou, C. Panayiotou, Foaming of polymers with supercritical CO2: An experimental and theoretical study, Polymer, 48 (2007) 59285939. [60] E. Kiran, Foaming strategies for bioabsorbable polymers in supercritical fluid mixtures. Part I. Miscibility and foaming of poly(l-lactic acid) in carbon dioxide + acetone binary fluid mixtures, Journal of Supercritical Fluids, 54 (2010) 308-319. [61] M. Xanthos, Functional Fillers for Plastics, Wiley, 2010. [62] S.A.E. Boyer, J.P.E. Grolier, Modification of the glass transitions of polymers by high-pressure gas solubility, Pure and Applied Chemistry, 77 (2005) 593-603. [63] C. Gualandi, L.J. White, L. Chen, R.A. Gross, K.M. Shakesheff, S.M. Howdle, M. Scandola, Scaffold for tissue engineering fabricated by non-isothermal supercritical carbon dioxide foaming of a highly crystalline polyester, Acta Biomaterialia, 6 (2010) 130-136. [64] E. Kiran, Foaming strategies for bioabsorbable polymers in supercritical fluid mixtures. Part II. Foaming of poly(ε-caprolactone-co-lactide) in carbon dioxide and carbon dioxide + acetone fluid mixtures and formation of tubular foams via solution extrusion, Journal of Supercritical Fluids, 54 (2010) 296-307. [65] M. Karimi, M. Heuchel, T. Weigel, M. Schossig, D. Hofmann, A. Lendlein, Formation and size distribution of pores in poly(ε-caprolactone) foams prepared by pressure quenching using supercritical CO 2, Journal of Supercritical Fluids, 61 (2012) 175-190. [66] E. Kiran, Modification of biomedical polymers in dense fluids. Miscibility and foaming of poly(p-dioxanone) in carbon dioxide + acetone fluid mixtures, Journal of Supercritical Fluids, 66 (2012) 372-379. [67] I.A. Ronova, O.V. Sinitsyna, S.S. Abramchuk, A.Y. Nikolaev, S. Chisca, I. Sava, M. Bruma, Study of porous structure of polyimide films resulting by using various methods, Journal of Supercritical Fluids, 70 (2012) 146-155. [68] H.X. Huang, H.F. Xu, Preparation of microcellular polypropylene/polystyrene blend foams with tunable cell structure, Polymers for Advanced Technologies, 22 (2011) 822-829. 251 Supercritical Antisolvent and Foaming Processes [69] H. Zweifel, R.D. Maier, M. Schiller, Plastics Additives Handbook, Carl Hanser GmbH, 2009. [70] L.J.M. Jacobs, M.F. Kemmere, J.T.F. Keurentjes, Sustainable polymer foaming using high pressure carbon dioxide: A review on fundamentals, processes and applications, Green Chemistry, 10 (2008) 731-738. [71] K. Wasai, G. Kaptay, K. Mukai, N. Shinozaki, Modified classical homogeneous nucleation theory and a new minimum in free energy change. 1. A new minimum and Kelvin equation, Fluid Phase Equilibria, 254 (2007) 67-74. [72] K. Wasai, G. Kaptay, K. Mukai, N. Shinozaki, Modified classical homogeneous nucleation theory and a new minimum in free energy change. 2. Behavior of free energy change with a minimum calculated for various systems, Fluid Phase Equilibria, 255 (2007) 55-61. [73] K. Nishioka, I. Kusaka, Thermodynamic formulas of liquid phase nucleation from vapor in multicomponent systems, The Journal of Chemical Physics, 96 (1992) 5370-5376. [74] S. Cotugno, E. Di Maio, G. Mensitieri, S. Iannace, G.W. Roberts, R.G. Carbonell, H.B. Hopfenberg, Characterization of microcellular biodegradable polymeric foams produced from supercritical carbon dioxide solutions, Industrial and Engineering Chemistry Research, 44 (2005) 1795-1803. [75] Y.T. Shieh, K.H. Liu, The effect of carbonyl group on sorption of CO 2 in glassy polymers, Journal of Supercritical Fluids, 25 (2003) 261-268. [76] J. Rincón, P. Cañizares, M.T. García, Improvement of the waste-oil vacuumdistillation recycling by continuous extraction with dense propane, Industrial and Engineering Chemistry Research, 46 (2007) 266-272. [77] J. Wang, X. Cheng, X. Zheng, M. Yuan, J. He, Preparation and characterization of microcellular polystyrene/polystyrene ionomer blends with supercritical carbon dioxide, Journal of Polymer Science, Part B: Polymer Physics, 41 (2003) 368-377. [78] S.P. Nalawade, V.H.J. Nieborg, F. Picchioni, L.P.B.M. Janssen, Prediction of the viscosity reduction due to dissolved CO 2 of and an elementary approach in the supercritical CO2 assisted continuous particle production of a polyester resin, Powder Technology, 170 (2006) 143-152. [79] L.I. Cabezas, V. Fernández, R. Mazarro, I. Gracia, A. De Lucas, J.F. Rodríguez, Production of biodegradable porous scaffolds impregnated with indomethacin in supercritical CO2, Journal of Supercritical Fluids, 63 (2012) 155-160. [80] A.R. Imre, G. Melnichenko, W.A. Van Hook, Liquid-liquid equilibria in polystyrene solutions: The general pressure dependence, Physical Chemistry Chemical Physics, 1 (1999) 4287-4292. [81] T. Kalwarczyk, N. Ziebacz, S.A. Wieczorek, R. Holyst, Kinetics and dynamics of dissolution/mixing of a high-viscosity liquid phase in a low-viscosity solvent phase, Journal of Physical Chemistry B, 111 (2007) 11907-11914. 252 Chapter 8 Life Cycle and Economic Assessment Chapter 7 shows the foaming process of PS from its homogeneous solutions in terpene oils using CO2 as antisolvent and blowing agent. Once the phases’ equilibrium was studied and the effect of CO2 on the shift of physical properties was measured, the design of a foaming process was carried out. Initially, the feasibility of the foaming was determined on the selection of the vanishing points of the solutions in CO2. Next, a further study concerning the optimization of the process in order to achieve microcellular foams from PS/Limonene solutions. Life Cycle and Economic Assessment 254 Chapter 8 Graphical abstract 255 Life Cycle and Economic Assessment RESUMEN Normalmente, se cree que el reciclado de residuos plásticos es una de las alternativas de gestión que presenta mayores ventajas medioambientales. Sin embargo, también puede tener algunas implicaciones negativas, relacionadas con emisiones peligrosas, contaminación de suelos, o incluso riesgos para la salud. La identificación y evaluación de los impactos medioambientales sobre el ciclo de vida del proceso explicado en este trabajo, es el objetivo de este Capítulo. Así, se comprobó que la tecnología propuesta presenta numerosas ventajas respecto a las técnicas comúnmente usadas. Debido a los resultados tan prometedores que mostró el reciclado de residuos de Poliestireno, se estimó su viabilidad económica. En primer lugar, se determinó el coste de la inversión y se realizó un análisis financiero en dos escenarios distintos, en función del volumen de residuos tratados. A continuación se calculó el Valor Actual Neto (VAN) y la Tasa Interna de Rentabilidad (TIR), ya que estos valores se usan normalmente para determinar la inversión y la rentabilidad de un proceso. Por último, se obtuvo la matriz DAFO, que es una metodología de estudio de la situación de un proyecto, analizando sus características internas (Debilidades y Fortalezas) y su situación externa (Amenazas y Oportunidades) en una matriz cuadrada. 256 Chapter 8 ABSTRACT Generally, recycling of plastic wastes could be thought as an environmentally beneficial alternative respect to other management ways. Unfortunately, there could be negative implications due to hazardous emissions, soil contamination or health risks. The identification and evaluation of the environmental impacts over the life cycle of the process is the aim of the first section of this Chapter. Thus, it was checked that the proposed technology presents several environmental advantages versus the traditional techniques used. Due to the promising results offer by this process for the recycling of Polystyrene wastes, its economic feasibility was investigated. Initially, the investment and financial analysis were determined in two different scenarios depending on the amount of treated wastes. Next, the Net Price Value (NPV) and the Internal Rate of Return (IRR) were calculated since the mentioned parameters are generally used for determining investments and projects profitability. Finally, the StrengthsWeaknesses-Opportunities- Threats (SWOT) matrix of the process was determined, revealing the interest of the process for industrial application. 257 Life Cycle and Economic Assessment 258 Chapter 8 8.1. Life Cycle Assessment Background Life Cycle Assessment (LCA) is an environmental management tool which identifies all resources used and wastes generated to all environmental media (air, water and soil) over the whole life cycle of a specific good or service [1]. LCA was introduced in the early 1970s, by simple calculations of energy usage during the entire life span of the product [2]. The main asset of LCA is its comprehensive character which make it very popular in analyzing plastic wastes management as it is shown by the numerous published studies of the life cycle emissions of these systems [1]. The increasing number of research papers has been focused on the LCA, particularly concerning the recycling of plastics (Figure 8.1). 3000 Life Cycle Assessment Research Paper 2500 2000 1500 1000 500 0 1970 1975 1980 1985 1990 1995 2000 2005 2010 2015 Year 18 16 Life Cycle Assessment Plastics Recycling Research Paper 14 12 10 8 6 4 2 0 1994 1996 1998 2000 2002 2004 2006 2008 2010 2012 2014 Year Figure 8.1. Evolution of research papers concerning Life Cycle Assesment (a) and Life Cycle Assesment of plastics recycling (b) along the last decades. Recycling is shown as environmentally beneficial because it reduces resources, decreases demand for landfill space and promotes savings energy, in principle. Nevertheless, the collection of wastes also shows environmental impacts as a 259 Life Cycle and Economic Assessment consequence of collection, sorting and recovery. By this reason, it is necessary to evaluate the costs and benefits from the recycling as well as the emissions generated. One methodology for undertaking this evaluation is LCA, which quantifies the environmental impacts of a product or material over its entire lifecycle [3]. An LCA comprises four major stages: goal definition, inventory, impact assessment and improvement assessment. According to this classification, the section below has been divided according to the mentioned aspects. The aim of this research is the recycling of Polystyrene wastes according to an environmentally friendly technology. The new recycling process proposed in this work is shown in Figure 8.2. Customer Shrinking and Transport Recycling Plant High Quality Polystyrene Volume reduction: 1/5-1/35 Figure 8.2. Schematic diagram of the recycling process of Polystyrene wastes for the production of microcellular foams. Figure 8.2 shows a schematic diagram of the recycling methodology where the PS wastes from customer are dissolved in Limonene and transported to the recycling plant by the trucks. Recycled microcellular foam PS can be reused as raw material for high quality applications and recovered Limonene can be reused to shrink new PS wastes. Next, the inventory is defined, mass and energy balances required to produce the functional unit are established and the environmental burdens are quantified [4]. The different studied scenarios were: 1. 2. 3. 4. 5. 6. Virgin Polystyrene production Landfilling Pyrolysis of Polystyrene wastes Thermal treatment of Polystyrene wastes Dissolution of Polystyrene wastes in Limonene Dissolution of Polystyrene wastes in Limonene, precipitation and foaming by scCO2. They are summarized in Figure 8.3. 260 Chapter 8 Low quality Recycled Polystyrene Emissions Emissions to to air air Emissions Emissions to to air air Emissions Emissions to to air air Landfill Landfill Transport Transport Emissions Emissions to to air air and and water water Polymerization Styrene Styrene Pyrolysys Pyrolysys Sorting Sorting and and Classification Classification Virgin Virgin Polystyrene Polystyrene Polystyrene Polystyrene Wastes Wastes Thermal Thermal treatment treatment Emissions Emissions to to air air Dissolution Dissolution in in terpene terpene oils oils Terpene Terpene oils oils Transport Transport Additives Additives for for Cyanoacrylates Cyanoacrylates CO CO22 Microcellular Microcellular Foams Foams Terpene Terpene oils/ oils/ CO CO22 High quality recycled Polystyrene Figure 8.3. Process flow of the different scenarios for the management of Polystyrene wastes. The different scenarios are shown below together with a brief description. 1. Virgin Polystyrene production Polystyrene is a commodity thermoplastic with many special properties which make it very useful in a wide range of applications. Styrene monomers are polymerized via chain-growth by free-radicals initiation with the aim to produce atactic Polystyrene (aPS). Once the radical initiator initiates (generally, benzoyl peroxide) the polymerization of styrene, propagation occurs which “builds up” the polymer chain. Due to the reactivity of vinyl monomers, it is often necessary to add an inhibitor to stabilize the monomer and prevent premature radical formation and polymerization. Once the polymer chain has “grown” and at a desirable length or molecular weight, the polymerization is terminated [5]. The polymer is then isolated, possibly purified, characterized, and used for material use. On the other hand, syndiotactic Polystyrene (sPS) is polymerized by coordination using metallocenes catalysts [6]. sPS, in contrast to aPS is semi-crystalline, and exhibits enhanced chemical resistance and thermal stability. Although the improved properties of sPS, aPS is more widely ended up in waste streams and by this reason, the LCA of aPS polymerization was studied. According to Figure 8.3, during the polymerization of styrene, emissions to the air and water are produced. 261 Life Cycle and Economic Assessment 2. Landfilling The amount of space used in landfills by all plastics is between 25-30 %. Although Polystyrene wastes account for less than 1% of the total weight of landfill materials, but the fraction of landfill space it takes up is much higher due to their low density. Furthermore, it is essentially non-biodegradable, taking hundreds or thousands of years to decompose. Even after disposal in landfill, EPS can easily be carried by the wind to litter streets or pollute water bodies or could be eaten by animals. Moreover, toxic chemicals could leach out of the wastes into the rainwater and be moved into the soils. From Figure 8.2 it is observed that the main emissions of PS wastes in landfills are to the air. 3. Pyrolysis of Polystyrene wastes The pyrolysis process for plastics takes the long chain polymer molecules and breaks or cracks them into shorter chains through heat producing energy. Energy recovery is an effective way to reduce the volume of organic materials by incineration but this method has been widely accused of being ecologically unacceptable, owing to the health risk from air born toxic substances [7]. Depending on the components of the waste plastics used as feedstock for energy recovery, the resulting emissions may contain contaminants such as amines, alcohols, waxy hydrocarbons and some inorganic substances [8]. Due to the wide variety of pollutants emitted during the energy recovery, the main substances produced during the pyrolysis of real PS wastes were analyzed and results are shown in Figure 8.4. 100 3.0 90 2.5 70 2.0 60 50 1.5 40 1.0 30 20 (a) Weight Deriv. Weight 10 Deriv. Weight (%/ºC) Weight (%) 80 C2H2 -HCN CH3Cl 0.5 0 0.0 C2H2 Intensity (a.u.) (b) Intensity (a.u.) -HCN CH3Cl C4H3 + C4H4 + C4H3 + C4H4 + 74 C6H6 C8H8 74 C6H6 C8H8\ 50 100 150 200 250 300 350 400 Temperature (ºC) 450 500 550 350 400 450 500 Temperature (ºC) Figure 8.4. Thermogravimetric and differential thermogravimetric analysis for the pyrolysis of PS wastes (a) and main components produced during its pyrolysis (b). 262 Chapter 8 According to Figure 8.4 during the pyrolysis of PS wastes highly toxic pollutants are produced due to the degradation of the polymer. The main free radical generated is C4H3+, but also cyclohexane and benzene are yielded. 4. Thermal treatment of Polystyrene wastes PS wastes become in densified PS by compacting at temperature above its Tg ( > 100 ºC). Typically, an EPS cushion package consists of 2% Polystyrene and 98% air. Removal of the air by mechanical or thermal densification results in a weight-tovolume ratio 30 to 50 times greater than the raw EPS package. The PS pellets cannot be recycled as high quality polymers due to the thermal degradation and ageing that suffer along the process [9, 10]. Densification will tend to lock in contaminates, lowering or negating the value of the densified materials. Moreover, the melting of the polymer requires high energy consumption and high temperature could promote the release of harmful vapours. 5. Dissolution of Polystyrene wastes in Limonene The dissolution of PS wastes has been shown not only as a volume reduction way, but also as a potentially cheap source for further applications. In this work, its application as fillers in cyanoacrylates based adhesives is studied. The addition of PS/Limonene solutions in different cyanoacrylate monomers was studied. Limonene was added as carrier in a variable concentration in order to homogenize the dissolution of PS into cyanoacrylates and avoid their instantaneous polymerization. PS was added as filler at maximum concentration of 9% (from a solution at 19% in Limonene) while the adhesive properties of cyanoacrylates remained constant or still enhanced. From the application of the explained management of PS wastes only emissions to air due to the transport of solutions should be considered for the study of the life cycle assessment. 6. Dissolution of Polystyrene wastes in Limonene, precipitation and foaming by scCO2. The process described in Chapter 7 allows the recovery of a high quality PS from wastes by an environmentally friendly recycling process. According to Figure 8.3, during the process any emissions are produced to air or water since all streams are recirculated. By this reason, the environmental impact should only consider the transport of PS wastes and results will be comparable to the previous section. There are many references in the literature concerning the Life Cycle Assessment of the synthesis of polymers, landfilling, recycling process and energy recovery [11-13]. In this section, the impact assessment was limited to the main resources consumed, the CO2, SO2 and NOx emissions and other environmentally relevant factors, such as acidification or greenhouse effect. Figure 8.5 summarizes the mentioned factors. 263 Life Cycle and Economic Assessment CO2 emissions SO2 emissions Landfill of PS Virgin PS NOx emissions CO2 emissions SO2 emissions NOx emissions Greenhouse effect Acidification Energy 0 10 20 30 40 50 60 70 Emissions to air 80 0 10 20 30 40 50 60 70 40 50 60 70 Emissions to air Thermal treatment CO2 emissions SO2 emissions Pyrolysis NOx emissions CO2 emissions SO2 emissions NOx emissions Greenhouse effect Acidification Energy 10 15 20 25 30 0 10 20 30 Emissions to air Foaming 5 Emissions to air Dissolution 0 CO2 emissions SO2 emissions NOx emissions CO2 emissions SO2 emissions NOx emissions Greenhouse effect Acidification Energy 0 3 6 9 12 Emissions to air Greenhouse effect Acidification Energy 15 0 3 6 9 12 Emissions to air 15 Figure 8.5. Amount of emissions to air from 1 kg of PS. The amount of CO2 emissions to air from 1 kg of PS in the different scenario is shown in Figure 8.5. The most important CO 2 emissions to the atmosphere are due to the landfill and pyrolysis of PS wastes. Otherwise, the lowest CO 2 emissions are achieved by the thermal treatment, dissolution and foaming of PS which produces between three and twelve times lower than that from landfill or pyrolysis. Furthermore it is important to highlight that during PS wastes collection and dissolution, the volume reduction enables a more efficient transportation due to the volume reduction which reduces significantly the emissions during the transport. Regarding SO2 emissions, the pyrolysis and landfill of PS are the main sources of emissions, followed by the thermal treatment while the dissolution of wastes in terpenes emits the lowest value of SO2. On the other hand, NOx emissions show the 264 Chapter 8 same trend as that of SO2. This fact is due to the emission related to the efficient in transport and by this reason the production of virgin PS exhibits low SO 2 and NOx values of emissions because transport is not included, moreover the higher density of PS pellets means lower volume of transport than the processed wastes. Finally, other emissions such as CH4, dioxins or CO should be considered, but due to the lack of data makes it quite difficult. Other parameters to evaluate the environment impact of the management of PS wastes were normalized and included. Thus, the greenhouse effect, the acidification potential and the equivalent energy consumption were quantified. According to data shown in Figure 8.5 the most negative impacts are due the production of virgin polymer, while the rest of alternatives are more environmentally friendly, especially, those related with recycling and recovery of PS wastes. In view of these facts, the results of this LCA indicate that recycling scenarios are generally the environment preferable options versus all impact categories. Unfortunately, the different references show an important disparity and due to the high sensitivity of the impacts, results could vary considerably. Nevertheless, if high quality recycled PS (as additive fillers of microcellular foams) is produced and it can substitute virgin plastic, these are the most favourable alternatives for the waste management of PS wastes. According to the explained facts, the economical assessment of the recycling of PS wastes for the production of high quality products, microcellular foams, is shown in the subsequent section. 265 Life Cycle and Economic Assessment 8.2. Economical evaluation One of the main objectives of a plant feasibility study is to assess the commercial viability of the proposed plant. The essential parameters of this calculation are the capital cost and cost of debt, the operating costs, the revenue on operation, and the plant life [14]. On a regular basis, supercritical technology presents high investment requirements due to the high pressure operation [15] but the products obtained after the supercritical process present improved characteristics which make them high added value products [16]. Among the most promising purposes of supercritical fluid processes are materials and pollution abatement both areas are covered in this work. Furthermore, the recycling of Polystyrene wastes entails huge environmental benefits which would be sufficient to convince consumers to change their habits and customs on polymer applications. In this section, the economical assessment includes the calculation of the equipment costs required for the process and then the total capital investment of the plant. According to the promising results shown in Chapter 7, the recycling of Polystyrene to produce microcellular foams would be the most beneficial alternative. Two different scenarios have been fixed to accomplish the calculations and to compare the feasibility of the process based on market demand. 1.- Building and operating a new plant located in Alcázar de San Juan due to the potential management capacities which offers the village. The scope is the recycling of PS wastes from Castilla-La Mancha region. In this scenario, the considered amount of PS wastes was 7000 kg/day according to the population of the region and the rates of the mentioned polymer wastes produced per habitant in Spain [17]. Along the following sections, the amount of treated wastes will be discussed as a function of the price. 2.- Construction and operating trucks containing the facilities for the recycling of daily PS wastes. The considered area is also Castilla-La Mancha region. In the case of scenario number 2, due to the smaller capacities of the plant (restricted to the volume of the truck), the treated PS wastes were 100 kg/day since it is the estimated value from household yielding. In the case of the industrial plant (scenario 1), five vessels for the foaming of PS wastes were designed because they allow to avoid dead times, reduce the energy requirements and fulfil the recycling of 7000 kg/day of Polystyrene wastes [18]. However, when the recycling process is carried out in the trucks (scenario 2), only one vessel was designed due to volumetric restrictions. The flow diagram of the plant (scenario 1) is shown in Figure 8.6, where the operating pressure and temperature of the main equipments are reported. PS wastes are accumulated in different collection points and dissolved in tank trucks containing Limonene. The solution and the CO 2 are fed in liquid and gas phase by a membrane pump (P-1) and a compressor (K-1), respectively to their corresponding 266 Chapter 8 foaming tank (C-1). In the foaming tanks (C-1 to C-5), Polystyrene is precipitated adopting microcellular structures and once CO 2/Limonene vapour phase is removed by means of the corresponding valves, it is discharged by means of a conveyor belt (C-8). The saturated CO2/Limonene solution is expanded by a turbine (K-2) and in the separator tank (C-6) both Limonene and CO2 can be recovered in liquid and gas phase, respectively. CO2 is returned to the feed tank (C-7) by means of a compressor (K-3) while collection trucks are fed by Limonene in order to dissolve again new PS wastes. The study of the economical feasibility of CO2 recovery (using compressor K3) was performed and results are shown below. A set of heat exchangers were designed and optimized with the aim of recovery a 99.99% of Limonene during the depressurization of the mixture CO 2/Limonene. The selection of the most suitable temperature to achieve the condensation of the terpene was performed according to the equilibrium results shown in Chapter 4. Finally to improve energy efficiency of the process and take advantage from the thermal shift, the heating or cooling water forms part of a closed circuit, where it is recirculated. On the other hand, for the recycling of PS wastes in recovery trucks, the flow diagram was modified to be fitted into them. The modified flow diagram of the recycling plant for the scenario 2 is represented in Figure 8.7. In this case, only one foaming vessel was designed and plant requirements were also modified. In this case, PS wastes are directly fed to the foaming vessel (C-1) that is previously filled with the Limonene where dissolution and foaming are carried out in two steps, but in the same vessel. CO2 is pressurized by the compressor (K-1) and fed to the foaming vessel. The process is performed following the previously described procedure. Due to the lower amount of PS wastes treated, the economical feasibility of the recirculation compressor (K-3) was not considered in scenario 2. On a regular basis, recovery of CO2 is only valuable when high volume of gas is required. Thus, an easier diagram is obtained, which will affect mainly to the investment costs. 267 -5 -57 E-2 5 5 30 1 25 30 Pressure (bar) Temperature (ºC) P T 25 C-0 C-7 1 E-1 P-1 84 K-1 90 90 5 I-1 C-1 I-2 C-2 I-3 C-3 I-4 C-4 I-5 C-5 20 K-2 C-8 E-3 50 217 50 K-3 C-6 PIC Life Cycle and Economic Assessment Figure 8.6. Schematic diagram of the industrial plant for the recycling of PS wastes to produce microcellular foams. 268 Chapter 8 I-24 C-3 C-2 I-20 K-1 K-2 C-1 C-4 Figure 8.7. Schematic diagram of the truck containing the recycling plant for the management of PS wastes to produce microcellular foams. 3.2. Investment analysis Next, equipments defined in the previous section were designed and their costs were determined to accomplish the first step of the investment analysis. Methods to estimate equipment cost are usually based on correlations which link a specific characteristic of the equipment with its final price. Generally, these correlations are referred to previous years and cost must be updated to the current year [19]. The accuracy of these methods depend on the specific type of equipment and the working conditions, therefore some authors think that the most accurate method for determine process equipment costs is directly from suppliers [20]. In this work, the correlations studied by Peters et al.[19] were used and prices were updated according to the overall Consumer Price Index (CPI) [21]. Equipments described in Figures 8.6 and 8.7 have been grouped in vessels, impulsion devices and heat exchangers. Tables 8.1-8.3 summarize the most relevant characteristics of the different equipments and their updated costs scenario 1. 269 Life Cycle and Economic Assessment Table 8.1. Estimated tanks and vessels cost (€2014) for the recycling of Polystyrene wastes plant in scenario 1. Equipment C-0 C-1 to C-5 C-6 C-7 Volume (m3) 2.1 15.83 0.42 Cost (€2014) 28844.3 347094.6 15428.9 100000 The impulsion devices of scenario 1 are compressors (K-1 and K-3), pump (P-1) and turbine (K-2). Table 8.3. Estimated impulsion equipments cost (€2014) for the recycling of Polystyrene wastes plant in scenario 1. Equipment K-1 K-2 K-3 P-1 W (kW) 2020 -1969 2714 17.2 Cost (€2014) 774887.5 739175.3 7860550.9 34935.0 Next, the optimization of the compressor (K-3) will be studied because it implies high investment costs. Heat exchangers are necessary to keep constant the temperature not only in the vessels but also in the pipes and valves in order to optimize the global process. Nevertheless, during compression important heat is produced that can be useful to balance the temperature decrease during expansions. The integration of energy is especially highlighted in scenario 1 because of the high values of flow. Furthermore, the design of the heat exchanger was similar to simplify the construction and building. Table 8.3. Estimated heat exchangers cost (€2014) for the recycling of Polystyrene wastes plant in scenario 1. Equipment E-1 E-2 E-3 A (m2) 60.32 60.32 60.32 Cost (€2014) 18374.1 18374.1 18374.1 Table 8.4 summarizes the cost of the equipments. All of them were designed in stainless steel 316 and the thickness of the vessels was calculated according to the most restrictive criteria. 270 Chapter 8 Table 8.4.Estimated equipment cost (€ ) for PS wastes recycling plant in scenario 1. 2014 Equipment C-0 C-1 to C-5 C-6 C-7 K-1 K-2 K-3 P-1 E-1 E-2 E-3 Cost (€2014) 28844.3 347094.6 15428.9 100000 774887.5 739175.3 7860550.9 34935.0 18374.1 18374.1 18374.1 TOTAL EQUIPMENT COSTS 11,344,417.20 Capital investment and operating costs The capital investment represents both the amount of money which must be supplied for the manufacturing and plant facilities, namely fixed capital investment, and that required for the operation of the plant, known as working capital. Percentage of delivered equipment cost is the method used for estimating the fixed or total capital investment. The other items included in the total direct plant cost are then estimated as percentage of the delivered equipment. The addition components of the capital investment are based on average percentage of total direct plant cost total direct and indirect plant costs or total capital investment. Estimating by percentage of delivered equipment cost is commonly used for preliminary and study estimates [19, 22]. Using the fixed-percentage method the capital investment was calculated and results are shown in Table 8.5. Table 8.5. Capital investment (€2014) by the percentage of delivered-equipment cost in scenario 1. Equipments installation 53,31,876.1 Instrumentation 2,041,995.1 Piping 7,487,315.4 Electrical services 1,247,885.9 Buildings 2,041,995.1 Land improvement 1,134,441.7 Utilities services 7,941,092.0 Land DIRECT COSTS 680,665.0 39,251,683.5 271 Life Cycle and Economic Assessment Table 8.5. (Continued) Capital investment (€2014) by the percentage of delivered-equipment cost in scenario 1. Engineering 3,743,657.68 Construction 4,651,211.05 Taxes 453,776.69 Contractor Fees 2,495,771.78 Contingency 4,991,543.57 INDIRECT COSTS 16,335,960.77 Operating costs include the manufacturing operation, the raw materials, and energetic requirements, utilities and plant maintenance. The cost of raw materials was not directly taken into account because PS wastes are provided freely by the consumers, but the transportation costs from the collection point to the recycling plant were considered as 0.7 €/km [23]. In both scenarios, a working radius of 200 km/truck was defined to cover the area of influence. Energetic requirements were calculated from the electrical power used for pumping and compression; the price of electricity was established at 0.117 €/kWh as it was shown in Eurostat [24]. With the aim of decrease energy consumption, the energy produced during the expansion of the CO2/Limonene stream in K-2 was used to feed K-1 and K-3. Maintenance and repairs were established as 7% of the total capital investment attending to the complexity of the process due to the high values of operating conditions, especially of pressure, and taxes were calculated as 0.5% of the fixed capital investment. The labour costs were established according to the number of workers in each category: truck drivers, operators and engineer. The summary of the operating costs are shown in Table 8.6. Table 8.6. Operating costs for the recycling of PS wastes process in scenario 1. Raw materials Electricity Direct operating labour Maintenance and repairs Directive management Taxes 260,470.6 1,253,743.0 210,000.0 1,667,629.3 52,500.0 277,938.2 Total 425,8048.0 Income statement Finally, the Net Price Value (NPV) is generally used for judging investments and projects, it means, whether the company made or lost money during the period being reported. NPV is the difference between the present value of cash inflows and the present value of cash outflows. NPV analysis is sensitive to the reliability of 272 Chapter 8 future cash inflows that an investment or project will yield and is used in capital budgeting to assess the profitability of an investment or project. If the NPV of a prospective project is positive, the project should be accepted. However, if NPV is negative, the project should probably be rejected because cash flows will also be negative. On the other hand, Internal Rate of Return (IRR) is the discount rate often used in capital budgeting that makes the NPV of all cash flows from a particular project equal to zero. Generally speaking, the higher a project's internal rate of return, the more desirable it is to undertake the project. As such, IRR can be used to rank several prospective projects a firm is considering. Assuming all other factors are equal among the various projects, the project with the highest IRR would probably be considered the best and undertaken first. To determine the optimum selling price and to consider the trend market, the Polystyrene price curve for the recycling process was determined in the range of production between 700 and 9000 kg/day as shown in Figure 8.8. 60 55 50 45 Price (€/kg) 40 35 30 25 20 15 10 5 0 0 1500 3000 4500 6000 7500 9000 Sales of recycled Polystyrene (kg/day) Figure 8.8. Price curve for the recycling of Polystyrene wastes into microcellular foams by dissolution and supercritical CO2 in scenario 1. Each marker of the curve shown in Figure 8.8 represents for a specific sales, the price to meet the Break-even-Point, it means, the point at which a business begins to make profits. According to Figure 8.8., when the sales of the recycled Polystyrene increase, its price decreases. Considering the amount of wastes in Castilla-La Mancha region and assuming that all the wastes are recycled (7000 kg/day), a sale price of 4.02 €/kg assures the economic feasibility of the explained process. Nevertheless, a higher price was considered (6 €/kg) to achieve economical benefits from the recycling process. 273 Life Cycle and Economic Assessment The estimation of the income statement for the recycling of Polystyrene wastes considered a linear amortization in 15 years (6.67%). Considerations for the accounting period 2014–2030 are reported in Table 8.7. Table 8.7. Income statement over the accounting period 2014-2030. Recycling of PS wastes in scenario 1. Amortisation along 15 years Investment capital Inflation Working capital 0.045 Investment curve 55,587,644.3 82,298.1 Sales (€) 12,600,000.0 3,722,281.2 t año=0 0.60 Costs (€) t año=1 0.40 Amortisation 0.067 Interest rate 10% Taxes 0.35 Figure 8.9 represents the cash flow and the Net Price Value along the temporary horizon of the recycling plant in scenario 1. Cash Flow (€) Net Cash Flow (€) 10000000 0 Cash Flow (€) 2014 2016 2018 2020 2022 2024 2026 2028 2030 Temporary horizon -10000000 -20000000 -30000000 -40000000 Figure 8.9. Evolution of Cash Flow and Net Cash Flow along the operating time of the recycling plant in scenario 1. Next, the most relevant results for the recycling of Polystyrene wastes in scenario 1 are shown in Table 8.8. 274 Chapter 8 Table 8.8. Results from the cash flow analysis for the recycling of PS wastes. Temporary horizon: 2 years of proyect+15 years of operation Price of recycled PS (€/kg) 6 Price of recycled PS (€/dm3) 1.82 NPV > 0 6,584,829.0 € IRR (%) 11.73 From these studies, it could be concluded that proposed process seemed to be an interesting process to recycled Polystyrene wastes into high added-value foam since the final product is industrially interesting due to the wide range of applications that it can cover at a low price. Next, the study of feasibility of recovering CO2 was performed. In this case, CO2 would not be recycled to the storage tank, and consequently, the compressor K-3 would not be necessary; however, CO2 should be supplied to the recycling plant in each batch. The price curve increases significantly as it is observed in Figure 8.10. 800 700 Price (€/kg) 600 500 400 300 200 100 0 0 1500 3000 4500 6000 7500 9000 Sales of recycled Polystyrene (kg/day) Figure 8.9. Price curve for the recycling of Polystyrene wastes into microcellular foams by dissolution and supercritical CO2 in scenario 1 without CO2 recovery. From the comparison of Figure 8.8 and 8.9 is observed that recovery of CO 2 decreases more than order of magnitude the cost of the recycled Polystyrene. Considering 7000 kg/day as the treated amount of wastes, a sale price of 75.42 €/kg assures the economic feasibility of the explained process. Nevertheless, a higher price was considered (76 €/kg) to achieve economical benefits from the recycling process. Thus, the main results concerning the feasibility of the explained option are shown in Table 8.9 275 Life Cycle and Economic Assessment Table 8.9. Results from the cash flow analysis for the recycling of PS wastes without CO 2 recovery. Temporary horizon: 2 years of proyect+15 years of operation Price of recycled PS (€/kg) 76 Price of recycled PS (€/dm3) 23.11 NPV > 0 1,463,869.9 € IRR (%) 11.22 According to Table 8.9, the CO2 recovery is more advantageous than the continuous supply of the gas and by this reason, this option was discarded for the scenario 2. Following the explained procedure, the economical feasibility for the recycling of PS wastes in scenario 2 are shown below, where the amount of treated polymer is 100 kg/day. The cost of the vessels depicted in Figure 8.7 is shown in Table 8.10. Table 8.10. Estimated tanks and vessels cost (€2014) for the recycling of Polystyrene wastes plant in scenario 2. Equipment C-1 C-2 C-3 Volume (m3) 5.65 0.16 Cost (€2014) 71,349.1 10,514.1 100,000 The cost of the equipments in scenario 2 is very much affected by the smaller size of the installation. In the case of scenario 2, there were only one compressor (K-1) and a turbine (K-2) since the Limonene is recovered directly to the vessel C-1 while CO2 recovery was not economically valuable. The costs of impulsion equipments are shown in Table 8.11. Table 8.11. Estimated impulsion equipments cost (€2014) for the recycling of Polystyrene wastes plant in scenario 2. Equipment K-1 K-2 W (kW) 485.6 -1112.9 Cost (€2014) 733306.5 308308.4 It should be remarked that turbine (K-2) releases higher amount of energy than the required in the compression of CO2 and consequently it can be used in further application. Next, the Table 8.12 summarizes the cost of the main equipments in scenario 2. All of them were designed in stainless steel 316 and the thickness of the vessels was calculated according to the most restrictive criteria. 276 Chapter 8 Table 8.12.Estimated equipment cost (€ ) for PS wastes recycling plant in scenario 2. 2014 Equipment C-1 C-2 C-3 K-1 K-2 TOTAL EQUIPMENT COSTS Cost (€2014) 71,349.1 10,514.1 100,000 733,306.5 308,308.4 1,223,478.1 The capital investment was calculated according to the percentage method [19, 22], but some budget items were missed since they are not necessary (for instance, land or buildings). The capital investment results are shown in Table 8.13. Table 8.13. Capital investment (€2014) by the percentage of delivered-equipment cost in scenario 2. Equipments installation 575,034.7 Instrumentation 220,226.1 Piping 807,495.5 Electrical services 134,582.6 Utilities services 856,434.7 DIRECT COSTS 3,817,251.67 Engineering 403,747.8 Taxes 48,939.1 Contractor Fees 269,165.2 Contingency 538,330.4 INDIRECT COSTS 1,260,182.44 As it was explained previously, operating costs include the manufacturing operation, the raw materials, and energetic requirements, utilities and plant maintenance and they are shown in Table 8.14. Table 8.14. Operating costs for the recycling of PS wastes process in scenario 2. Raw materials Direct operating labour Maintenance and repairs Directive management Taxes 253,994.1 40,000.0 152,323.02 10,000 25,387.17 Total 481,704.29 277 Life Cycle and Economic Assessment Before examining the feasibility of recycling Polystyrene wastes in trucks, the price curve as a function of the amount of wastes was studied and it is shown in Figure 8.10. 350 300 Price (€/kg) 250 200 150 100 50 0 0 20 40 60 80 100 120 140 160 180 200 220 Sales of recycled Polystyrene (kg/day) Figure 8.10. Price curve for the recycling of Polystyrene wastes into microcellular foams by dissolution and supercritical CO2 in scenario 2. According to Figure 8.10, the price of the recycled Polystyrene decreases when their sales increase. Setting the values of the treated polymer in 100 kg/day, a minimum sale price of 30.76 €/kg should be economically feasible. Nevertheless, it is important to highlight that if the amount of wastes increases up to 200 kg/day, the sale price decreases up to 15.36 €/kg, that presents the same order of magnitude than that obtained in the scenario 1.In order to achieve economical benefits from the recycling process of 100 kg of wastes per day, a higher sale price was set (35 €/kg). The income statement was determined in the same way that in the previous section. The most relevant parameters are shown in Table 8.15. Table 8.15. Income statement over the accounting period 2014-2030. Recycling of PS wastes in scenario 1. Amortisation along 15 years Investment capital Inflation Working capital 0.045 Investment curve 278 5,077,434.1 5500.0 Sales (€) 600,000.0 217,358.5 t año=0 0.60 Costs (€) t año=1 0.40 Amortisation 0.067 Interest rate 10% Taxes 0.35 Chapter 8 Figure 8.11 represents the cash flow and the Net Price Value along the temporary horizon of the recycling plant in scenario 2. 1500000 Cash Flow (€) Net Cash Flow (€) 1000000 500000 0 Cash Flow (€) 2014 2016 -500000 2018 2020 2022 2024 2026 2028 2030 Temporary horizon -1000000 -1500000 -2000000 -2500000 -3000000 -3500000 Figure 8.11. Evolution of Cash Flow and Net Cash Flow along the operating time of the recycling plant in scenario 2. Next, the most relevant results for the recycling of Polystyrene wastes in scenario 2 are shown in Table 8.16. Table 8.16. Results from the cash flow analysis for the recycling of PS wastes in scenario 2. Temporary horizon: 2 years of proyect+15 years of operation Price of recycled PS (€/kg) 30 Price of recycled PS (€/dm3) 9.12 NPV > 0 733,383.8 € IRR (%) 12.09 Table 8.16 shows that the design process could be also applicable to smaller volume of production according to the modified scheme, but considering a slightly higher sale price. Although the recycling process of Polystyrene by means of an industrial plant or by recycling trucks might be a challenge, but the subsequent considerations should be taken into account: There could be government subsidies and aid to promote the recycling of the polymer. 279 Life Cycle and Economic Assessment Any management costs were considered, which could decrease the costs of the process. Future bans concerning the use of Polystyrene are expected. According to the mentioned points, the recycling of Polystyrene wastes by dissolution in terpenes and CO2 at high pressure for the production of microcellular foams is stated as a promising technology. Thus, the studied recycling process is presented as environmentally and economically advantageous since it is valuable even though the aims and management profits were not considered in the economical analysis. In summary, the recycling of Polystyrene wastes for the production of microcellular foams could be positioned as very valuable in an immediate future. Finally, the analysis of SWOT matrix (Table 8.17) will be a helpful tool to determine and clarify the real interest of a process, discarding non interesting ideas for industrial applications. Table 8.17. SWOT matrix for the recycling of Polystyrene by dissolution and supercritical CO2 to produce microcellular foams. Strengths Environment friendly Clean facilities Recycling that reduces overall volume as well as generates profits Low resource consumption Cheap raw material Previous experience on scCO2 Easy scale-up and transport of the plant Favourable access to distribution network Established organizational structure Opportunities Employment opportunities New technology No competence on recycling XPS/EPS with high quality New regulation about waste management Social conscience about responsible consumption Unfulfilled customer need Strategic positioning Trade Monopoly Weaknesses Finite storage facilities Poor participation by resident in materials collection High cost plant High cost process No previous experience on trade Only one scale-up plant Weak brand name Trained and qualified staff Threats High capital requirements New competitors Emergence of similar products or technology Social and economical instability Low level of financial resources As can be seen in Table 8.17, the analysis of opportunities and strengths confirmed the interest of the microcellular foams process. Nowadays, the environmental 280 Chapter 8 awareness is increasing but the consumption of Polystyrene does not decrease, which assure a potential market of the mentioned polymer products. Furthermore, the wide knowledge of the system provides an important versatility respect to the target products, which could change attending the necessity of the markets. It means, not only microcellular foams can be produced, but also, higher concentration solutions without PS ageing, or in a future, microparticles and foams. From the analysis of the threats and weaknesses of Table 8.17, the high cost of the new technology could be overcome if customers are able to differentiate the new product from regular Polystyrene, which should be focused on highlighting the environmentally benefits of the new products. 281 Life Cycle and Economic Assessment References [1] J. Cleary, Life cycle assessments of municipal solid waste management systems: A comparative analysis of selected peer-reviewed literature, Environment International, 35 (2009) 1256-1266. [2] B. Hannon, SYSTEM ENERGY AND RECYLING: A STUDY OF THE CONTAINER INDUSTRY, ASME Pap, (1972). [3] A.L. Craighill, J.C. Powell, Lifecycle assessment and economic evaluation of recycling: A case study, Resources, Conservation and Recycling, 17 (1996) 75-96. [4] A. Azapagic, A. Emsley, I. Hamerton, Polymers: The Environment and Sustainable Development, Wiley, Guildford, UK, 2003. [5] P.J. Flory, Principles of Polymer Chemistry, Cornell University Press, 1953. [6] J. Schellenberg, Syndiotactic Polystyrene: Synthesis, Characterization, Processing, and Applications, Wiley, 2009. [7] D.S. Achilias, I. Kanellopoulou, P. Megalokonomos, E. Antonakou, A.A. Lappas, Chemical recycling of polystyrene by pyrolysis: Potential use of the liquid product for the reproduction of polymer, Macromolecular Materials and Engineering, 292 (2007) 923-934. [8] F. Pinto, P. Costa, I. Gulyurtlu, I. Cabrita, Pyrolysis of plastic wastes. 1. Effect of plastic waste composition on product yield, Journal of Analytical and Applied Pyrolysis, 51 (1999) 39-55. [9] M. Guaita, THERMAL DEGRADATION OF POLYSTYRENE, British Polymer Journal, 18 (1985) 226-230. [10] A. Karaduman, E.H. Imşek, B. Çiçek, A.Y. Bilgesü, Thermal degradation of polystyrene wastes in various solvents, Journal of Analytical and Applied Pyrolysis, 62 (2002) 273-280. [11] T. Noguchi, H. Tomita, K. Satake, H. Watanabe, A new recycling system for expanded polystyrene using a natural solvent. Part 3. Life cycle assessment, Packaging Technology and Science, 11 (1998) 39-44. [12] R.B.H. Tan, H.H. Khoo, Life cycle assessment of EPS and CPB inserts: Design considerations and end of life scenarios, Journal of Environmental Management, 74 (2005) 195-205. [13] S. Madival, R. Auras, S.P. Singh, R. Narayan, Assessment of the environmental profile of PLA, PET and PS clamshell containers using LCA methodology, Journal of Cleaner Production, 17 (2009) 1183-1194. [14] P. Watermeyer, Handbook for Process Plant Project Engineers, John Wiley & Sons, 2002. [15] M. Perrut, Supercritical fluid applications: Industrial developments and economic issues, Industrial and Engineering Chemistry Research, 39 (2000) 45314535. [16] M.P. Fernández-Ronco, A. De Lucas, J.F. Rodríguez, M.T. García, I. Gracia, New considerations in the economic evaluation of supercritical processes: Separation of bioactive compounds from multicomponent mixtures, Journal of Supercritical Fluids, 79 (2013) 345-355. [17] ANAPE, http://www.anape.es/ in, Asociación Nacional del Poliestireno Expandido, 2014. [18] G.A. Núñez, C.A. Gelmi, J.M. del Valle, Simulation of a supercritical carbon dioxide extraction plant with three extraction vessels, Computers and Chemical Engineering, 35 (2011) 2687-2695. 282 Chapter 8 [19] M. Peters, K. Timmerhaus, R. West, Plant Design and Economics for Chemical Engineers, McGraw-Hill Education, 2003. [20] K. Hillstrom, L.C. Hillstrom, Encyclopedia of Small Business: A-I, Gale Group, 2002. [21] INE, http://www.ine.es/calcula/index.do?L=1, in, Instituto Nacional de Estadística, 2014, pp. Update a personal income or spending with the overall CPI. [22] R.H. Perry, D.O.N.W.A. GREEN, J.O.H. Maloney, Perry's Chemical Engineers' Handbook, McGraw-Hill Professional Publishing, 1997. [23] M.d. Fomento, www.fomento.gob.es, in, Madrid, 2014, pp. Observatorio de mercado del transporte de mercancías por carretera. [24] Eurstat, Electricity prices for industrial consumers, in, http://epp.eurostat.ec.europa.eu/tgm/table.do, 2013. 283 Life Cycle and Economic Assessment 284 Chapter 9 Conclusions and Future work Chapter 7 shows the foaming process of PS from its homogeneous solutions in terpene oils using CO2 as antisolvent and blowing agent. Once the phases’ equilibrium was studied and the effect of CO2 on the shift of physical properties was measured, the design of a foaming process was carried out. Initially, the feasibility of the foaming was determined on the selection of the vanishing points of the solutions in CO2. Next, a further study concerning the optimization of the process in order to achieve microcellular foams from PS/Limonene solutions. Conclusions and Future Work 286 Chapter 9 Fruto de los resultados obtenidos en esta investigación, se pueden concluir que: La obtención de productos de alto valor añadido, a partir de residuos de Poliestireno, mediante disolución en aceites naturales y separación con CO 2 a alta presión es viable. El proceso de disolución de residuos de Poliestireno en aceites terpénicos como paso inicial del proceso de reciclado es ventajoso, ya que el residuo presenta una elevada solubilidad en los disolventes mencionados, que se incrementa con la temperatura, sin presentar signos de degradación. Los resultados son extrapolables a residuos reales de Poliestireno Extruido y Expandido. La introducción de CO2 a alta presión en el sistema Poliestireno/aceite terpénico, provoca dos efectos principales. Por un lado, los terpenos presentan una elevada solubilidad en el CO2, en condiciones moderadas de presión y temperatura, por lo que son eliminados con relativa facilidad. Por otro lado, el CO2 se adsorbe en el polímero, provocando su plastificación y por lo tanto, facilitando su procesado. El proceso de separación de la disolución Poliestireno/aceite terpénico utilizando CO2 como antisolvente a alta presión, es técnicamente viable. La precipitación del polímero a partir de su disolución en terpenos, adicionando CO2 como antisolvente, de acuerdo con una estructura predefinida (micropartículas, fibras, espumas…) resulta muy compleja. Es necesario conocer el equilibrio de las mezclas ternarias, sus propiedades y la hidrodinámica del sistema. El proceso de espumación de disoluciones Poliestireno/aceite terpénico permite la separación del polímero adoptando una estructura microcelular. Las condiciones más adecuadas para llevar a cabo el proceso y conseguir espumas con un menor tamaño de poro y una distribución homogénea de los mismos, fueron alta presión (90 bar), baja temperatura (30 ºC), baja concentración (0.05 g PS/ml terpeno), elevado tiempo de contacto y despresurización (4 horas y 30 minutos, respectivamente). A partir del análisis económico y del ciclo de vida, se confirma la viabilidad del proceso de reciclaje de residuos de Poliestireno mediante disolución y CO2 a alta presión. Además de las conclusiones generales, los resultados más reseñables por Capítulos son: Capítulo 4: i) La estimación de la solubilidad del Poliestireno (PS) en disolventes de la familia de los terpenos a partir de métodos teóricos es una buena alternativa frente a las 287 Conclusions and Future Work tradicionales determinaciones experimentales. Así, se puede estudiar el efecto de la temperatura y el peso molecular sobre la solubilidad del polímero. ii) Los aceites terpénicos son totalmente solubles en CO 2 en condiciones moderadas de temperatura (25- 40 ºC) a presiones cercanas al punto crítico del gas, lo que permitiría su separación de las mezclas Poliestireno/aceite terpénico. iii) El CO2 se absorbe entre las cadenas poliméricas del Poliestireno causando el hinchamiento y la plastificación del polímero. Este fenómeno se observa a través de la modificación de sus propiedades a nivel macromolecular: descenso de la temperatura de transición vítrea (Tg), viscosidad () y tensión interfacial (). Capítulo 5: iv) El equilibrio de las mezclas ternarias formadas por CO 2/aceite terpénico/Poliestireno está principalmente determinado por la presión, la temperatura y la concentración de polímero en la disolución inicial. Un aumento de presión, favorece tanto la solubilidad del terpeno en la fase rica en CO2, como la adsorción del CO2 en la disolución. El incremento de la temperatura promueve un aumento de la solubilidad de los aceites terpénicos en CO 2, y también favorece la solubilidad del Poliestireno en el disolvente, por lo que es necesario estudiar cuál de los efectos es más relevante. Por último, cuando aumenta la concentración del polímero en la disolución, la solubilidad del aceite terpénico y la absorción del CO 2 disminuyen. Capítulo 6: v) El efecto de la presión sobre la viscosidad de las mezclas CO2/aceite terpénico/Poliestireno mostró que no sólo hay que tener en cuenta el efecto plastificante causado por el CO2, sino también el efecto de compresión que ejerce éste contra la disolución polimérica. El descenso de la viscosidad de las mezclas ternarias como resultado del incremento de la temperatura, sigue la ley de Arrhenius, a partir de cuyos resultados se puede calcular la energía de activación. Por último, el efecto más relevante que causa el aumento de la viscosidad es la adición de polímero a la disolución inicial. vi) La tensión interfacial de las disoluciones Poliestireno/terpeno disminuye linealmente cuando aumenta la presión de CO2 sobre la mezcla ternaria. Sin embargo, el efecto de la temperatura y la concentración no fueron muy significativos en el rango de estudio. vii) La temperatura de transición vítrea de las mezclas ternarias se ve principalmente afectada por el efecto plastificante del CO2 más que por la presencia del disolvente (terpeno). Capítulo 7: viii) La morfología del Poliestireno precipitado a partir de su disolución en aceites terpénicos y utilizando CO2 como antisolvente, depende principalmente de la 288 Chapter 9 transferencia de materia. A partir del cálculo de los coeficientes de difusividad y de la aplicación de la segunda ley de Fick, se puede conocer el perfil de concentraciones del CO2 y del terpeno en el interior de una gota de disolución. ix) La modificación de las propiedades del polímero en disolución favorecen la nucleación y el crecimiento de celdas, cuando se ponen en contacto CO2 y una disolución PS/aceite terpénico. Por este motivo, el proceso de espumación de disoluciones poliméricas permite la producción de espumas microcelulares de una manera relativamente sencilla. Capítulo 8: x) Las distintas alternativas de reciclaje de residuos de Poliestireno presentan ventajas medioambientales respecto al resto de técnicas de gestión de los mismos. En concreto, el reciclaje de residuos de Poliestireno mediante disolución en aceites terpénicos y espumación utilizando CO2 a alta presión, se posiciona como una de las mejores opciones para el tratamiento de residuos poliméricos, por ser medioambientalmente sostenible. xi) La evaluación económica del proceso estudiado, confirma que el reciclaje de residuos de Poliestireno para la producción de espumas microcelulares presenta mayores ventajas si el CO2 y el disolvente terpénico son recirculados en el proceso y el volumen de la planta de tratamiento es superior a 7000 kg/día. 289 Conclusions and Future Work From the present research about the following main conclusions can be concluded: The production of high-added value products from the recycling of Polystyrene wastes using natural terpene oils and high pressure CO 2 is feasible. The dissolution of Polystyrene wastes in terpene oils is the first step for the recycling since the polymer is highly soluble in terpenoids and its solubility is enhanced by the increase of temperature. During the dissolution process at mild temperatures, any degradation was observed. The results concerning solubility can be applied to real wastes, i.e. Extruded or Expanded Polystyrene. The addition of high pressure CO2 in the dissolution Polystyrene/terpene oils cause two effects. On one hand, the easy removal of terpenes from the mixture due to their high solubility in CO2 and could be easily removed. On the other hand, the plasticization of the polymer due to the CO 2 sorption which enhance the processing. The separation process of the Polystyrene/terpene oils using high-pressure CO2 as antisolvent is technically feasible. The precipitation of the polymer according to a previously selected morphology (microparticles, fibres, foams…) is not so evident. Equilibrium and hydrodynamics considerations are necessary to be studied. The foaming of Polystyrene/terpene oil solutions using CO2 is an easy process which allows the recovery of the polymer according to microcellular structures. The most suitable conditions to achieve foams with smaller, well defined and homogeneous cells are high pressure (90 bar), low temperature (30 ºC), low concentration (0.05 gPS/ml terpene), high contact time (4 hours) and high depressurization time (30 min). The life cycle and economic assessment showed excellent results concerning the feasibility of the explained recycling process for the Polystyrene wastes. In addition to the aforementioned conclusions, the most relevant results from the different chapters are: Chapter 4: i) The solubility of Polystyrene in terpene oils prediction is an acceptable alternative versus the traditional experimental measurements. The influence of the temperature and the polymer molecular weight on the solubility are studied. ii) Terpene oils are fully soluble in CO2 at mild temperatures (25- 40 ºC) and pressure close to the critical point of the gas, which allows the separation of binary mixtures composed by Polystyrene and terpene oils. 290 Chapter 9 iii) CO2 is absorbed among the polymer chains of Polystyrene causing its swelling and plasticisation. This fact is observed through the decrease of glass transition temperature (Tg), viscosity () and interfacial tension (). Chapter 5: iv) The equilibrium of ternary mixtures containing CO2/terpene oil/Polystyrene is mainly affected by pressure, temperature and concentration of polymer in the solution. When pressure increases, the solubility of terpenes in CO2 and the sorption of CO2 in the dissolution are enhanced. An increase of temperature enhances the solubility of terpenes in CO2, but also, promotes the solubility of Polystyrene in terpenes. Finally, when the concentration of polymer in the initial solution increases, the solubility of the terpenoids and the sorption of CO 2 decreases. Chapter 6: v) The effect of pressure on the viscosity of CO2/terpene oil/Polystyrene showed that not only plasticisation of solutions should be considered, but also the compression effect of the gas on the solutions. The decrease of viscosity due to the increase of temperature follows an Arrhenius-type equation and activation energy can be calculated. The most relevant effect on the increase of viscosity of the ternary mixtures is caused by the increase of polymer concentration. vi) Interfacial tension of Polystyrene/terpene oil solutions decreases linearly when pressure of CO2 in the ternary mixtures CO2/terpene oil/Polystyrene increases. The effect of temperature and concentration were not very significant in the studied range. vii) The glass transition temperature of the ternary mixture is mainly affected by the plasticising effect of CO2 instead of the presence of a solvent (terpene oil). Chapter 7: viii) The morphology of the recovered Polystyrene, from its solution in terpenes and using CO2 as antisolvent, is mainly dependant on the mass transfer. From the knowledge of the mutual diffusion coefficients and the application of Fick’s second law, the concentration profile was calculated. ix) The modification of the polymer properties enhances the nucleation and the cell growth during the foaming process. Thus, microcellular foams are easily produced. Chapter 8: x) Among the different alternatives for the plastic wastes management, the recycling was risen as the most environmentally friendly. Particularly, the recycling of Polystyrene wastes by dissolution in natural terpene oils and foaming using CO2 291 Conclusions and Future Work as blowing agent is arranged as one of the best options for the polymer wastes treatment. xi) The economic evaluation of the explained process confirm that the recycling of Polystyrene wastes for the production of microcellular foams takes more advantages if the CO2 and the terpenes are recirculated when the treated wastes are higher than 7000 kg/day. 292 Chapter 9 A partir de las conclusiones obtenidas en este trabajo de investigación, se recomienda: i) Desarrollar un modelo teórico completo y fiable que permita la predicción del comportamiento de mezclas ternarias a alta presión con el fin de ampliar el proceso desarrollado y aplicarlo a otros polímeros y disolventes. ii) Determinar las isopletas del sistema ternario y la influencia de la concentración de CO2 en la separación de las mezclas, lo que permitiría hacer una selección inicial de las condiciones más adecuadas para precipitar el polímero. iii) Llevar a cabo un estudio pormenorizado de la influencia de las condiciones hidrodinámicas en la precipitación de disoluciones formadas por Poliestireno y aceites terpénicos. iv) Diseñar y construir una instalación experimental para conseguir la precipitación de micropartículas de Poliestireno a partir de su disolución en terpenos y utilizando CO2 como antisolvente. v) Incorporar sustancias de alto valor añadido a las espumas microcelulares para aumentar la utilidad del producto final. vi) Estudiar el proceso de reciclaje de residuos plásticos mediante disolución y CO2 a alta presión. vii) Mezclar los residuos de Poliestireno con otros con un importante volumen de producción, como la celulosa. viii) Optimizar las condiciones de operación de la planta piloto con el fin de adaptarse a criterios económicos y a las demandas del mercado. 293 Conclusions and Future Work After the conclusions obtained in this research work, several guidelines for future work can be proffered: i) To develop a complete and reliable theoretical model for the prediction of the behaviour of ternary mixtures at high pressure in order to enlarge the application of the process to new polymers and solvents. ii) To determine the isoplets and the influence of CO2 concentration on the separation of ternary mixtures for a first screening of the most suitable conditions to precipitate the polymer. iii) To carry out a detailed study about the influence of hydrodynamic on the precipitation of Polystyrene/Terpene oils solutions. iv) To design and build an experimental setup for the efficient precipitation of Polystyrene microparticles from its solution in terpenes using CO2 as antisolvent. v) To load the microcellular foams with high-added value to increase its value. vi) To study the recycling of new plastic wastes by dissolution and high-pressure CO2. vii) To blend the Polystyrene with traditional wastes (cellulose). viii) To optimize the pilot plant operational conditions on the basis of economic criteria and market trends. 294 NOMENCLATURE Abbreviations: AAD: Average Absolute Deviation ABS: Poly (Acrylonitrile-co-butadiene-co-styrene) ANOVA: Analysis of Variance aPS: atactic Polystyrene ASES: Aerosol Solvent Extraction System ASD: Average Standard Deviation BPR: Back-Pressure Regulator CCD: Charged Coupled Device CFCs: Chlorofluorocarbons CNT: Classical Nucleation Theory CPI: Consumer Price Index DOE: Design of Experiments DSA: Drop Shape Analysis DSC: Differential Scanning Calorimetry DTGA: Differential Thermogravimetric Analysis EoS: Equation of State EPA: Environmental Protection Association EPS: Expanded Polystyrene FFEs: Fractional Factorial Experiments GAS: Gas Antisolvent GPC: Gel Permeation Chromatography HCFC: Hydrochlorofluorocarbon HDPE: High Density Polyethylene IFT: Interfacial Tension (mN/m) IRR: Internal Rate of Return LCA: Life Cycle Assessment LDPE: Low Density Polyethylene MHS: Mark Houwink and Sakurada empirical relationship 295 MW: molecular weight (g/mol) NPV: Net Price Value PC: Polycarbonate PCA: Precipitation with a Compressed Antisolvent PDI: Polydispersity Index PET: Polyethylene Terphthalate PI: Pressure indicator (manometer) PR-EoS: Peng-Robinson Equation of State PP: Polypropylene PS: Polystyrene PU: Polyurethane PVC: PolyVinyl Chloride RCS: Refrigerated Cooling Systems RESS: Rapid Expansion of Supercritical Solution SAS: Supercritical Antisolvent scCO2: supercritical CO2 SCF: Supercritical Fluid SEDS: Solution Enhanced Dispersion by Supercritical Fluids SEM: Scan Electron Microscope SL-EoS: Sánchez-Lacombe Equation of State sPS: syndiotactic Polystyrene Std. Dev.: Standard Deviation SWOT: Strenghts, Weakness, Opportunities and Threats Matrix TGA: Thermogravimetric Analysis THF: Tetrahydrofuran TIC: Temperature Indicator Controller VLE: Vapour-Liquid Equilibrium XPS: Extruded Polystyrene 296 Symbols: a: attractive forces between the molecules in Peng-Robinson Equation of State a: polymer chain rigidity applied to Mark-Houwink-Sakurada A: area of the micrograph (cm2) b: molecules size B: dimensionless factor depending on the viscosity of the material used in the Classical Nucleation Theory C’H: saturation of the cavities (g CO2/ g PS) C: concentration (g Polystyrene/ml terpene) C0: initial concentration (g/ml) Cpp: excess transition isobaric specific heat of the pure polymer (J/g·K) D: hydraulic diameter (m) Dab: diffusion coefficient of substance a in b (m2/s) erf: Gaussian error function Ea: flow activation energy (J/mol) g: acceleration of gravity (m2/s) Gm: Gibbs free energy (J/mol) H: Henry's law constant (g CO2/ bar·g PS) Hm: enthalpy of mixing (J/mol) Jss: number of critical nuclei formed during the Classical Nucleation Theory kB: Boltzmann’s constant (J/K) kH: analogous to Henry’s law used in Dual-Mode sorption model kij: coupling or binary interaction parameter of components i and j m: mass of the gas molecule in the Classical Nucleation Theory (g/mol) M: magnification factor Md: molar mass of the dissolved gas (g/mol) Mm: molar mass of the monomeric unit (g/mol) Mo: Morton dimensionless number n: number of cells N: number of gas molecules per unit volume (number of molecules/cm3) Nhom: rate of homogeneous nucleation (cells/s) 297 p: critical level of significance in the Analysis of Variance P: pressure (bar) reduced pressure P*: characteristic pressure (bar) Pc: critical pressure (bar) PR: reduced pressure Pv: vapour pressure (bar) P: difference between the pressure of the metastable and nucleating phase used in the Classical Nucleation Theory (bar) Q: flow (m3/s) r: size parameter used in Peng-Robinson Equation of State r: radius of the droplet used in Fick’s second law (m) R: ideal gas constant (J/K·mol) Re: Reynolds dimensionless number S: solubility (kg/m3) tcontact: contact time during foaming process (min) tdep: depressurization time during foaming process (min) T: temperature (ºC) : reduced temperature T*: characteristic temperature (ºC) Tc: critical temperature (ºC) Tb: boiling temperature (ºC) Tdeg: degradation temperature (ºC) Tg: glass transition temperature (ºC) Tg,0: glass transition temperature of the pure polymer at atmospheric pressure (ºC) Tm: melting temperature (ºC) : velocity of the liquid jet (m/s) V: molar volume (cm3/mol) xi: mole fraction of the component i in the liquid rich phase W: reversible work of critical nucleus formation (J) wi: mass fraction of the component i 298 wCO2Cym: total amount of CO2 retained in p-Cymene (g) wCO2PS: total amount of CO2 retained in the Polystyrene (g) wCO2sol: total amount of CO2 retained in the solution (g) yi: mole fraction of the component i in the vapour rich phase z: lattice coordination number 12: binary interaction parameter of terpene/polymer systems surface tension (N/m) d : solubility parameter from dispersion forces (MPa1/2) h: solubility parameters due to the hydrogen bonds between molecules (MPa1/2) i: global solubility parameter of the substance i (MPa1/2) p: solubility parameter from dipolar intermolecular forces (MPa1/2) mer interaction energy : viscosity (Pa·s) R (t): relative viscosity time dependant (Pa·s) (0): viscosity at atmospheric pressure(Pa·s) (t): viscosity at time t (Pa·s) (): viscosity at saturation (Pa·s) c: viscosity of the surrounding fluid (Pa·s) l: viscosity of the liquid (Pa·s) sol: viscosity of the PS/Limonene/CO2 solution (Pa·s) solvent: viscosity of the solvent (Pa·s) rel: relative viscosity between the solution and the solvent Ki: partition or distribution coefficients i: volume fraction of the substance i : density (kg/m3) c: density of the continuous phase (kg/m3) g: density of the gas phase (kg/m3) l: density of the liquid phase (kg/m3) : reduced density 299 : characteristic density (kg/m3) closest distance allowed between two mers (m) m: interfacial tension of the mixture PS/Terpene and CO 2 (mN/m) pol: interfacial tension of the polymer (mN/m) sol: interfacial tension of the solution of CO2 in PS (mN/m) v*: characteristic molar volume(cm3/mol) v0: hole volume (cm3/mol) 12: interaction enthalpy between terpenes and polymers (J/mol) : acentric factor Ω∞1: activity coefficient at infinite dilution of component i 300