JoVE Science Education
Chemical Engineering
A subscription to JoVE is required to view this content.  Sign in or start your free trial.
JoVE Science Education Chemical Engineering
Kinetics of Addition Polymerization to Polydimethylsiloxane
  • 00:07Overview
  • 01:18Principles of Polymerization
  • 03:48Polymer Synthesis
  • 06:06Results
  • 08:08Applications
  • 09:40Summary

Kinetik der Additionspolymerisation zu Polydimethylsiloxan

English

Share

Overview

Quelle: Kerry M. Dooley und Michael G. Benton, Department of Chemical Engineering, Louisiana Landesuniversität, Baton Rouge, LA

Polymere sind Moleküle bestehend aus vielen sich wiederholenden Monomer-Einheiten, die chemisch zu langen Ketten verbunden sind. Sie weisen eine Vielzahl von physikalischen Eigenschaften, die von ihrer chemischen Struktur, Molekulargewicht und Polymerisationsgrad betroffen sind. Die Polymer-Industrie produziert Tausende von Rohstoffen in einer breiten Vielzahl von kommerziellen Produkten. 1 , 2

Dieses Video soll eine Additionsreaktion Polymerisation führen und Werten Sie anschließend das resultierende Produkt um zu verstehen, wie Viskosität zur Polymers Molekulargewicht herangezogen werden kann. Darüber hinaus wird dieses Experiment untersuchen wie Molekulargewicht Monomer Umwandlung bezogen werden können.

Principles

Procedure

Das System wird gesteuert durch Steuersequenzen PS1-PS5 auf einen standard verteilten Industrieregler, der betrieben wird von einem PC ausgeführt. Die Sequenzen öffnen/schließen/anpassen Ventile in die richtige Reihenfolge und informieren, wann und wie Sie den Reaktor Komponenten hinzufügen. 1. Reaktor Aufbau Offene N2 Zylinder an den Reaktionsbehälter angeschlossen. Führen Sie die Steuersequenz PS1, die Ausrüstung zu testen. Schließen Sie dann …

Results

The molecular weight can be determined by empirical relationships, such as Barry's relationship for polydimethylsiloxanes with molecular weights above ~2,500.5

Equation 2

This gives the viscosity-average molecular weight. For molecular weight predictions < 2,500, interpolate the experimental data found in Kuo,6 using the kinematic viscosity of the DC-245 monomer for chain length 1. Divide the viscosity (cP) by the polymer density (g/cm3) to obtain the kinematic viscosity in cSt. Divide the viscosity average MWs by 1.6 (empirical factor for PDMS) to get the number average molecular weight, and divide this value by the monomer molecular weight to get the average chain length, (PN)avg, which includes the unreacted monomer.

To get the fraction conversion (fm), start with the mass balance for the average of PN (polymer only):

Equation 3  (1)

The left-hand side is the average of PN (polymer only) up to time t, where f = fm. But the average PN that you measure includes the monomer. To account for monomer in (PN)avg, recall that by definition:3-4

Equation 4

and therefore:

Equation 5 (2)

The average polymer Equation 6 and (PN)avg for the entire batch are almost equal at the last batch, where fm approaches 1. Compute fm for the last point using a mass balance and the amount of low boilers that was collected. Solve for Equation 6. For many addition polymerizations, Equation 6 is constant for the entire batch, which allows fm to be computed at all other times from Equation 2. Also, compute the equilibrium constant K (first-order reversible kinetics model) for the reaction by mass balance.

Once fm has been determined as a function of time, assume irreversible kinetics and determine the reaction order with respect to monomer. Use statistical analysis to determine the quality of the fits and the confidence limit on the rate constant kp. Determine the fit for first-order kinetics (expected from theory),3-4 and test if the two fits actually differ.

At similar conditions, others have reported a first-order rate constant of 10-3 s-1 for the DC-245 monomer, and a K > 60.

Figure 2
Figure 2. Typical polymerization results."DOP" = degree of polymerization. The MW's were computed from available data (see ref. 6) or Barry's equation (>2500).5

The workup of representative raw data is shown in Figure 2. These data are for the polymerization of Dow Corning DC-245 monomer. The reaction conditions were: 0.04 wt% catalyst solution, 12 wt% endblocker (modifier), 130 °C and 1 atm pressure. With a relatively large amount of endblocker used, the final degree of polymerization (DOP) was quite low.

In this experiment, 11.36 L of monomer were reacted, and only 15 mL low boilers were recovered, indicating that the data should follow irreversible kinetics. The fit to first-order (in monomer) kinetics is shown in Figure 3 below. The fraction conversions (f) were determined using Equations 1 and 2 with the assumption that the polymer produced is at constant chain length (PN). The resulting fit is reasonable, but not perfect. Slight deviations from the theoretically expected first-order kinetics can arise for several reasons such as diffusional effects, which is when the viscosity increases and the diffusivities decrease significantly. Two other reasons for deviations are suggested by the raw reaction temperature data (temperature oscillations affect the rate constant) and by small leaks that may be present in the pumps, reactor, and heat exchangers. If there are leaks, some O2 could get into the system and gradually inhibit the reaction.

Figure 3
Figure 3. Kinetics analysis. "F" is the 1st order function, the solution of the batch reactor mass balance for a 1st order irreversible reaction.

Applications and Summary

Polymer science provides many examples of the basic principles of chemical kinetics and reactor design. Simple rate expressions can describe fairly complex chemical processes, as in this experiment. Reactor system design must find the optimal reactor type (batch, stirred tank, plug flow, or hybrid) considering kinetics, capital costs, and molecular weight distribution. In particular, the last factor is usually the most important, because it largely defines the product. Depending on this factor alone the product can often range from a hard brittle solid to a rubber to a liquid. A bulk (no solvent) polymerization, such as the one performed in this experiment, has the advantage that subsequent processing to obtain a pure polymer is simple – just strip out the low boilers and filter out the neutralized catalyst. However, the disadvantage of bulk polymerization is that if one loses control of temperature (too high), even in a thermoneutral polymerization, other reactions will dominate and lead to "runaway", which is an uncontrolled exothermic reaction that may result in explosion. Polymerizations with higher heats of reaction are reacted either in solution, suspension (a continuous water phase is present, and the monomer is in droplet form), or in the gas phase.

The major takeaways from the experiment are how one can process raw data of an easily measurable physical property (viscosity) to ultimately determine the monomer fraction conversions and the kinetics of the reaction. Many other physical properties, e.g., density and particle light scattering, are used for this purpose in other polymerizations.

Polymers made by ring-opening polymerizations include Nylon-6 from caprolactam, acetal copolymers with ethylene oxide and dioxolane, which are used in everything from fuel tanks to sprinklers, poly(ethyleneimines), which are used in detergents and cosmetics, and many other silicon-backbone polymers.With the exception of Nylon-6, most of these polymers are commercially produced by anionic or cationic polymerizations. Other polymers that are made similarly include copolymers of styrene (especially with isoprene), isobutene-isoprene (butyl) rubber and its halogenated variants, and poly(alkyl vinyl ethers), which are typically used in paints and adhesives. For some such polymerizations, the chain terminations are so controlled that an almost homogeneous molecular-weight distribution is possible. Except for certain specialty grades, it has been found that such narrow distributions present other problems, such as extrusion difficulties.

Many polymers are vacuum-stripped as the first part of their purification to a commercial product. Among these are the poly(vinylidene chloride) copolymers, poly(chloroprene), and many grades of poly(styrene) and its copolymers such as SAN (styrene-acrylonitrile).

Silicone polymers are used in many products, including lubricants, personal care products, medical devices, antifoams, sealants, waterproof coatings, and as components of detergents, electrical insulation, and paints.8 Medical devices composed of very high molecular weight crosslinked silicone may be approved by the FDA for implantation. More common medical uses are consumables such as catheters, tubing, gastric bags, and surgical incision drains. Commercial PDMS is non-hazardous with a flash point higher than 300 °C, minimal toxicological effects, and good resistance to moderately concentrated aqueous alkali and acids.8,9 It does not corrode most common materials. But like many polymers it can oxidatively decompose, in this case above ~150 °C.

Materials List

Name Company Catalog Number Comments
Equipment
Rotational (cup and bob) viscometer  Brookfield Use to measure the viscosity of polymer samples
Stirred tank reactor custom 20 L
Reactor Agitator

McMaster-Carr 46-460 RPM; 6-blade, flat turbine (Rushton) type, ~4 diameter.
Reagents
Dimethlysiloxane monomer Dow Corning DC-245 specific gravity = 0.956 at 25 °C; viscosity = 4.2 cSt; m = average number of dimethylsiloxanes = 5
Endblock A Dow Corning 10082-147 specific gravity = 0.88 at 25°C; m = 4.5 (not counting the two end groups)
KOH catalyst VWR 470302-140 45 wt% solution in water
Nitrogen Airgas UHP grade Used to blanket the system
Carbon dioxide Airgas Tech. grade Used to neutralize the catalyst

Viscosity and Density Data at Low Molecular Weight

Data originally from: Dow Corning.10

MW, g/mol 162 410 1250 28000
Viscosity, cs, 25 °C 0.65 2.0 10 1000
Specific gravity, 25 °C 0.760 0.872 0.935 0.970

References

  1. http://www.essentialchemicalindustry.org/polymers/polymers-an-overview.html and http://www.pslc.ws/mactest/maindir.htm (both accessed 8/22/16).
  2. MatWeb, Material Property Data, http://www.matweb.com/ and Plastics General Polymers Brand Name Listing, http://www.plasticsgeneral.com/BRAND-NAMES-LIST1.htm (both accessed 8/25/16).
  3. Fogler, F.S., Elements of Chemical Reaction Engineering, 3rd Ed., Prentice-Hall, 2001, pp. 354-382 (sections 7.1.2-7.1.5).
  4. Odian, G., Principles of Polymerization, 4th Ed., Wiley-VCH, 2004 (ch. 3), or Rodriguez, F., Principles of Polymers Systems, 2nd Ed., McGraw-Hill, 1982 (ch. 4); Fried, J.R., Polymer Science and Technology, Prentice-Hall, 1995 (ch. 2).
  5. Barry, A.J., Viscometric Investigation of Dimethylsiloxane Polymers, J. Appl. Phys., 1946, 17, 1020-1024.
  6. Kuo, A.C.M. Poly(dimethylsiloxane), in Polymer Data Handbook, Oxford University Press, 1999, 411.
  7. Ullmann’s Encyclopedia of Industrial Chemistry, Wiley-VCH, Weinheim, 2012 or the Encyclopedia of Polymer Science and Technology, 3rd Ed., Wiley-Interscience, Hoboken, 2003-04.
  8. http://www.dowcorning.com/content/discover/discoverchem/properties.aspx (accessed 8/25/16)
  9. Shin-Etsu Silicone Fluid Technical Data, Shin-Etsu Silicones of America, Akron, 2005.
  10. Dow Corning, Product Information, Silicon Fluids, http://www.dowcorning.com/applications/Product_Finder/Products.aspx  (accessed 9/23/16).

Transcript

Polymers are a ubiquitous class of compounds found in all facets of industry and manufacturing. Two of their most important characteristics, molecular weight and degree of polymerization, must be derived from other bulk properties. Unlike other substances, whose physical characteristics are defined solely by their chemical structures, polymers are also affected by their degree of polymerization and molecular weight. Chemically identical polymers can vary from liquids to rubbers to hard, brittle solids, all based on these physical properties. Since microscopic attributes, such as molecular weight, are difficult to measure directly, bulk properties, such as viscosity, density, and light scattering, can be used to infer these important characteristics. This video will illustrate a batch polymerization of polydimethylsiloxane, or PDMS, and determine its molecular weight and degree of polymerization from its viscosity.

To begin, let’s focus on the bulk fabrication of polydimethylsiloxane, or PDMS. Polymerization reactions are classified by their mechanisms, reactor types, product characteristics, and more. In the case of PDMS, an initiator reacts with the monomer to produce the polymer chain, which in turn can be extended through further reactions with the monomer. This reaction mechanism is known as addition polymerization, and is characterized by the absence of by-products. The choice of reactor depends on reactant properties and affects the characteristics of the product. Batch reactors, which typically consist of a tank, agitator, and heating or cooling system, operate as closed systems in which the reactants are added in a discreet step and then allowed to react over time. Batch reactors are preferred for small-scale reactions when low quantities of reactants are used or a new process is being developed, or to synthesize several grades of product. They are frequently used for polymerizations. PDMS is synthesized from a monomer, an initiator, and an end-blocker without any solvent, a condition known as bulk polymerization. The absence of solvents simplifies polymer processing, since the by-products and catalyst are easily separated from the polymer. However, the temperature must be carefully controlled, as with a water cooling jacket, to prevent exothermic runaway that may result in an explosion. Regardless of the reaction conditions, the measured physical properties of the product, such as the viscosity, are used to estimate the number-average molecular weight and weight-average molecular weight. Dividing the number average molecular weight by the molecular mass of the monomer yields the average chain length or degree of polymerization, which is related to conversion and reaction order. Now that you know the basics of polymerization, let’s see how to operate a small-scale batch reaction of PDMS and determine the reaction kinetics.

To begin the procedure, open the nitrogen cylinder connected to the reaction vessel. Run the first sequence, which verifies that the equipment is operational and in good working order. Next, test the system for leaks by closing the manual valve to the vacuum pump. Wait five minutes and verify that the pressure rise does not exceed 600 millimeters of mercury. Reopen the valve to remove any remaining atmosphere. Finally, close the manual valve and fill the system with nitrogen. The third module of the program adds the cyclic monomer to the reactor. The lower quantity ingredients, the catalyst and end blocker, are added through a small funnel called the adder tank. The reactor is now full and ready for polymerization. Start the fourth process and monitor the temperature. Once it rises above 105 degrees, begin collecting liquid samples from the sample draw point. Collect aliquots at intervals of at least every eight minutes. To know when the polymerization reaches equilibrium, monitor the power usage of the agitator. Once power has stopped increasing, the reaction is complete. At this point, open the carbon dioxide tank and valve and press the reaction complete push button to neutralize the catalyst. To begin the stripping sequence, open the manual valve to the vacuum pump and allow it to run for 15 minutes at a higher temperature. At this point, select stripping complete and collect the low boilers from the reaction into a flask. Allow the automated cool down process to run. Using the manufacturer’s instructions, measure the collected samples with a rotational viscometer. If the speed is set too high, no reading will be obtained and a lower speed will be chosen. These values will be used to determine the molecular weight distribution of the polymer.

A lot of information can be obtained from the relatively simple viscosity measurement. Dividing the viscosity of the PDMS sample by its density yields its kinematic viscosity. Empirical equations, such as Barrie’s relationship, relate kinematic viscosity to the viscosity-average molecular weight. Dividing the viscosity-average molecular weight by 1.6, another empirical factor for PDMS, yields the number-average molecular weight, the average weight per polymer chain. Dividing this by the weight of the monomer yields the average chain length or degree of polymerization, the number of monomer units in the polymer. However, since the calculated chain length includes the un-reacted monomer, it will be artificially low. A correction that accounts for the fractional conversion must be applied. Here are typical results for the viscosity-average molecular weight and degree of PDMS polymerization with reaction time. In this reaction, a large amount of end blocker, which stops chain growth and forms a trimethyl end group, was used, resulting in a low final degree of polymerization. The fractional conversion can also be determined as a function of time. By assuming irreversible kinetics and that the polymer was produced at a constant chain length, the reaction order with respect to monomer was determined to be first order, as confirmed by the reasonable fit. A rate constant of 0.054 inverse minutes was calculated, which agrees with other studies that report a first order rate constant of 0.06 inverse minutes for this monomer under similar conditions.

Synthetic polymers are found in a wide range of products, both on the industrial and commercial scale. Let’s look at a few common examples. Siloxane polymers, such as PDMS, can be industrially formed via several techniques, such as injection molding. They are suitable for diverse applications, including lubricants, sealants, detergents, electrical insulation, paints, and medical devices. Medical implants and probes, such as this prototype, are of particular note, as PDMS is non-hazardous, has minimal toxicological effects, and resists moderately concentrated acids and bases. For these reasons, the FDA has approved the use of PDMS in the medical field. PDMS synthesis is an example of ring-opening polymerization, a common form of chain-growth polymerization. In ring-opening polymerizations, the chain iteratively opens cyclic monomers to form successive reactive centers on the polymer. Depending on the system, the reactive center can be radical, anionic, or cationic. This process allows strict control of molecular weight distribution, though this can, in turn, cause issues with extrusion. It has been shown that having some higher molecular weight polymer in the mixture provides a more uniform extrudate.

You’ve just watched Jove’s introduction to addition polymerization. You should now understand the concepts of both polymerization and how viscosity can determine monomer conversion and kinetics. Thanks for watching.

Tags

Cite This
JoVE Science Education Database. JoVE Science Education. Kinetics of Addition Polymerization to Polydimethylsiloxane. JoVE, Cambridge, MA, (2023).