Вы находитесь на странице: 1из 14

Elimination of Amine Emissions from Polyurethane Foams: Challenges and Opportunities

F.M. Casati, J.M. Sonney, H. Mispreuve, A. Fanget, R. Herrington, J. Tu

Reprinted from The API 2001 Proceedings with permission from the Alliance for the Polyurethanes Industry.

Elimination of amine emissions from polyurethane foams: challenges and opportunities

F.M. CASATI, J.M. SONNEY, H. MISPREUVE, A. FANGET Dow Europe S.A International Development Center 13, rue de Veyrot - P.O. 3 CH-1217 Meyrin 2 (GE) Switzerland

R. HERRINGTON, J. TU The Dow Chemical Company 2301 North Brazosport Freeport TX 77541 U.S.A.

ABSTRACT Polyurethanes are a unique class of plastic in the sense that independent manufacturers buy liquid polyols and isocyanates and then react them in the presence of catalysts, surfactants and other additives in widely varying processing schemes to make the final commercial product. In some manufacturing scenarios, there is a growing concern about worker exposure to volatile organic chemicals (VOC) such as the amine catalysts. Emissions from manufactured articles that contain polyurethane components are also a concern. The polyurethane industry is continuing its legacy of successfully reducing the amount of volatile organic chemicals released from its operations. Examples of this include the total elimination of the classical chlorofluorocarbon blowing agent from flexible foams and the replacement of volatile antioxidants with less migratory ones. Today, many of the classes of ingredients commonly used in preparing polyurethanes are available in low-fogging or low VOC grades. The next big challenge being addressed by the industry is the elimination of amine catalyst vapors. This is important for both VOC and odor concerns. The catalyst suppliers have responded by introducing a variety of new, non-fugitive (sometimes called reactive) amine catalyst compositions. Their success at reducing amine emissions while providing catalytic activity is widely reported in the literature. Difficulties associated with reactivity profile adjustment, higher usage levels and losses in certain foam physical properties suggest that an improved level of technology is still needed. This paper will present a review of work The Dow Chemical Company is doing to develop flexible foam grades of polyether

polyol that help reduce and in some cases eliminate catalysts while improving process. Comparative data will be presented to illustrate the utility of these new polyols in helping the polyurethane industry continue to reduce volatile emissions. INTRODUCTION Flexible polyurethane foams are well-established industrial products with important technical and economic advantages. These materials have found their way into our daily lives through their use in the endless array of consumer products. To complement this wide-ranging success, significant development efforts continue today to improve targeted performance features like comfort, durability, resistance to moisture and temperature [1-3]. Improvements in the environmental impact of these products are also being gained through reductions in the amount of VOCs released during their manufacture. In some manufacturing scenarios, there is a concern about worker exposure to the vapors of traditional amine catalysts. Amine emissions from manufactured articles containing polyurethane components is a growing debate due to potential fogging, PVC (PolyVinyl Chloride) staining and PC (Poly Carbonate) degradation [4-6]. The industrial hygiene and safety issues related to handling and use of amine catalysts in polyurethane production are complex. It begins with the need to purchase one or more of these additives in separate containers that must be offloaded, warehoused, transported to the formulation room and in many cases physically handled by workers assigned the task of blending up a formulation. Most amine catalysts have exposure and odor properties such that personal protective equipment is required [7]. Empty working and shipment vessels need to be properly handled and prepared for disposal or shipment back to the vendor.

Due to the exothermic nature of the reactions and the extra heat applied in some curing scenarios, freshly prepared polyurethane products are always at an elevated temperature. With the classical non-reactive type amines, odorous vapors are released and in the absence of proper ventilation, worker exposure can occur [8-10]. When captured by an exhaust system, these amine vapors can represent a significant contribution to the total VOCs being released by the manufacturing plant. Even after cooling down and undergoing fabrication into finished products, trace amine odors are sometimes still detectable. Amine catalyst vapors have been identified as one component of the new car smell [11]. The catalyst suppliers have responded by introducing a variety of new, non-fugitive (sometimes called reactive) amine catalyst compositions. Examples of their success at reducing emissions while providing catalytic activity are recorded in the literature [12-18]. Difficulties associated with reactivity profile adjustment, higher usage levels and losses in certain foam physical properties suggest that an improved level of technology is needed. In this paper, we continue to report the previously announced activities at Dow to develop active polyether polyols [19]. Comparative data will be presented to illustrate the utility of these new polyols in helping the polyurethane industry continue to reduce volatile emissions. BACKGROUND Most commercially manufactured polyurethane foams are made with the aid of at least one catalyst. Many classes of compounds have been investigated and the amines and the organo-metallics have been found most useful. Various combinations of catalysts are used in order to establish a balance between the chain propagation or gelling (isocyanate with polyol) reaction and the blowing reaction (isocyanate with water). Any given catalyst molecule will normally have some activity for both reactions and it is common to find synergistic performance when using two or more catalysts in a formulation. The polymer formation rate and the gas evolution rate must be balanced so that the gas is entrapped efficiently in the gelling polymer while the foam cells develop sufficient strength to maintain their structure without collapse or shrinkage. Catalysts are also important for assuring completeness of reaction or cure in the finished foam. Some of the challenges to success in designing a catalyst package for a polyurethane production reaction are listed in Table 1.

Table 1. Challenges to catalyze polyurethane reactions Solubility parameters (including water) Steric hindrance Difference of NCO reactivity between isomers Acidity Viscosity build-up Effect of temperature on catalyst efficiency

The first requirement for good catalytic activity is that the catalyst molecule be highly compatible with the mixture of reaction raw materials. Molecular structure of the catalyst and the resulting steric hindrance will impact how easily the molecule can approach the reactants and facilitate the desired reactions [20]. The catalyst package needs to be effective when isocyanates of varying isomer content and reactivity are used. Trace acidity, common in some raw materials, impact the performance of the catalyst(s). Molecular mobility of the catalyst should be high so that it can continue to exert its effect as the system reacts and builds viscosity. Since the reactions are exothermic, the catalytic effect should not be overly influenced by a rise in temperature. All of these requirements are more easily met with the conventional, fugitive type amines rather than the newer reactive type amines. To address these challenges, the reactive amines are normally used at significant higher concentrations, hence sometimes are not fully bound in the polymer network [21]. In addition, most of the reactive amines are mono-functional and can be considered as polymer chain terminators (which generally detracts from the ultimate properties of the polymer). Dow active polyols are useful to overcome these challenges. EXPERIMENTAL A list of all raw materials used with this study is given in Tables 2, 3 and 4 (SPECFLEX *, VORANOL *, VORANATE *, ISONATE * are trademarks of The Dow Chemical Company).
Table 2. List of raw materials SPECFLEX NC 630 Polyether polyol SPECFLEX NC 632 VORANOL CP 6001 VORANOL CP 1421 VORANOL 3137 A SPECFLEX NC 700 SAN copolymer polyol SPECFLEX NE 150 MDI isocyanate VORANATE T-80 TDI isocyanate

Table 3. List of catalysts


Trademark DABCO NE-200 DABCO NE-1060 DABCO 33 LV DABCO BLV DABCO VP 137 DABCO T TOYOCAT RX-20 TOYOCAT RX-21 PC-CAT HPI DMAMP-80 NIAX A-1 JEFFCAT ZF-10 Supplier Air Products Tosoh Nitroil Angus Crompton Huntsman

Molded Foam Processing Tests The protocol for testing foam processability factors have been described in detail [22] and will be further discussed below. A variety of heated aluminum molds of various shapes were used to evaluate the processing characteristics of both control and experimental foam formulations. Foam Testing Foam physical properties were determined in accordance with ASTM, ISO, NF and major OEMs (Original Equipment Manufacturers) testing procedures. PVC Staining Tests Accelerated aging tests at elevated temperature were carried out in closed containers in the presence of a PVC foil. A foam sample of 50x 50 x 25 mm was cut from the core of the foam part which had been isolated just after production and was placed at the bottom of a 1 liter glass jar. A piece of gray PVC skin, reference E 025 373A0175A, supplied by Bennecke-Kaliko (Germany), was hung by the rim of the jar, which was then sealed. Aging was carried out at 115C for 72 hours. After cooling, the PVC sheet discoloration was measured using a Minolta Chroma Meter CR 210. The smaller the change in color, the lower the E measured in this test. RESULTS AND DISCUSSION Dow newly developed active polyols were tested in various conditions with different technologies. The purpose of this work was to assess their performance and to define how they can be formulated to reduce and in some cases eliminate amine catalyst use. Initial work had been concentrated on the performance of reactive amines offered by the various suppliers and on their combination with Dows active polyols. In most cases it was found that these products gave good compatibility with Dow active polyols and that, by allowing a reduction of their concentration in polyurethane formulations, overall processing was improved. Main results of this study are presented hereafter with various technical options. These are concentrated on flexible foams since these are the focus of this research program. Each foam technology will be reviewed and discussed separately. SLABSTOCK FOAM Dows experience with slabstock foam has shown that most common density range can be produced without the addition of traditional amine catalyst. Experimental Polyol A can be used alone to provide a smooth rise profile for low density (16kg/m) formulations using all-water blowing. Rise rate of such formulation is identical to that of traditional amine catalyst blown formulations (Figure 1). Addi-

(DABCO* is a trademark of Air Products & Chemicals Inc; TOYOCAT * is a trademark of Tosoh Corporation; NIAX * is a trademark of Crompton Corporation; JEFFCAT * is a trademark of The Hunstman Corporation; PC-Cat * is a trademark of Nitroil Europe Handels GMBH; DMAMP-80* is a trademark of Angus Chemical Company) Table 4. List of Dow Experimental Active Polyols
Polyol A Additive A Polyol B Polyol C OH number 56 56 32 32 Functionality 3 <2 >3 >3 EO capping no yes yes yes

Foam Preparation Foam formulations are indicated with each section of the paper. Bench scale molded foams were produced in a 30x30x10 cm test block mold heated at 60C using the standard hand-mix procedure [22]. Free-rise foams were poured in a card-board box. Slabstock foams were either produced on a POLYMECH or on a VARIMAX machines. For MDI HR molding high pressure impingement mixing machines (KRAUSSMAFFEI and CANNON) were used to pour foams in a 40x40x10 cm standard test block mold and in a variety of production seat molds. Foams for the processability studies in low-density TDIbased HR molded foams were prepared in a Hi-Tech Engineering Model RCM-40 high pressure, impingement mixing foam machine. The machine was retrofitted with an Admiral brand, L-type (10/14) mixing head. Two process streams (isocyanate and polyol masterbatch) were used. Nominal throughput of the machine was 10 kg/min (22 pounds/minute) at the component ratios necessary to maintain a 105 isocyanate index for all foams made during this study.

tive A was used, together with Polyol A in formulations, especially those with low water level, where more blowing is necessary. Figure 2 illustrates rise rate comparison of 3.7 PHP (parts per hundred parts of polyol) water formulations with the new active Polyol A. With tuning of additive A, a rise rate similar to that of traditional amine catalyst can be reached.
6.5 Water Formulation
14

Table 5. Slabstock foam: Formulation and foam physical properties


Foam formulations
TDI INDEX VORANOL 3137A Exp.POLYOL A Exp ADDITIVE A WATER DABCO DC-5160 DABCO BLV DABCO T-9 6.5 1.2 0.12 0.15 0.15 6.5 1.2 3.7 1 0.15 0.21 0.21

A
115 100

B
115

C
110 100

D
110

100

99.3 0.7 3.7 1

12

10

8 Rise Rate 0.12Dabco BLV Exp. Polyol A 6

Reactivity
Cream Time (sec) Rise Time (sec) 13 90 13 91 12 101 12 103

Foam physical properties


CORE DENSITY (kg/m3) 90% COMPRESSION SETS CD (%)
0 20 40 60 Second 80 100 120 140

16.8 9.7 227 68 85

17.6 11.2 227 66 91

26.9 2.4 315 107 90

26.6 2.8 315 109 92

TEAR STRENGTH (N/m) ELONGATION @ BREAK (%) TENSILE STRENTH (Kpa)

Figure 1. Rise profile of 6.5 water formulation


Rise Rate , 3.7 Water Formulation
14

IFD HARDNESS (N) Mean Load at 25% Deflection Mean Load at 65% Deflection 201 381 110 1.89 21 211 399 114 1.89 32 203 366 141 1.80 41 193 354 135 1.83 42

12

Mean Load at 25% Return


10

Mean Modulus RESILIENCY (%)


Dabco BLV 0.15 pphp Exp. Polyol A + Add. A 0.7 pphp

8 Rise Rate

0 0 20 40 60 Second 80 100 120 140

Figure 2. Rise profile of 3.7 water formulation

Foams made with experimental Polyol A were tested for PVC staining after production in order to detect any amine emission. No staining was observed. In conclusion, it is feasible to consider that a slabstock foamer may in the future produce foam by just mixing polyols, water, surfactant and isocyanate. MDI HR MOLDING An emphasis has recently been put on elimination of amine emission in MDI based HR molding while maintaining good humid aging resistance. Severe tests, such as BMWs, VW-Audis, Volvos, Opel-GMs are already challenging when using conventional catalysis, as it has been described in previous articles [1,23]. The situation becomes more difficult when using reactive amine catalysts for the following reasons: 1) The catalyst level has to be increased in the formulation to maintain equivalent reactivity. Since these reactive catalysts are tied up in the polymer, they tend to disrupt it somewhat. 2) The gellation profile is changed when strong, fugitive catalysts such as TEDA (Triethylenediamine) are replaced with reactive amines. Hence the polymer build-up is slowed down or at least made differently. 3) With the high concentration of tertiary amines in the polymer chain there is always a possibility of retroreaction, or chemical degradation, upon foam exposure to a combination of heat and moisture. Indeed polymer mor-

Foams made with these new experimental polyols have identical physical properties with that of traditional amine blown foams, as can be seen in Table 5.

phology changes induced by accelerated aging are already reported with conventional foam systems [24]. The present work was focused onto the use of Active Polyol B together with reduced amount of conventional amines and/or reactive amines in order to meet VW-AUDI s technical (T2524-03) and environmental (PVC staining) requirements. This program is still underway at the time of writing this paper. Initial results presented in Table 6 confirm that HACSs (Humid Aged Compression Set) values of < 15 % are easily met and PVC staining is minimized. More challenging is the HALL (Humid Aged Load Loss), or CFD (Compression Force Deflection) change at 40 % deflection after humid aging, specifications of (-25 < HALL % < + 10) when fugitive amines are removed. Foam density of the tested foams was purposely reduced vs the 52 kg/m3 target to assess the limits.
Table 6. MDI HR molding Formulation
VORANOL CP 6001 SPECFLEX NC632 VORANOL CP 1421 Exp. Polyol B Water Niax A-1 Dabco 33-LV Catalyst A Catalyst B Diethanolamine LF Tegostab B-8715 LF SPECFLEX NE-150 Index 2 3.7 0.12 0.35 A 64 34 2 3.7 0.12 0.35 A 64 34 2 49 3.7 0.06 0.27 2 49 3.7 0.06 0.27 2 3.7 B 49 B 49 C 64 34 C 64 34 2 3.7

T-80 HR MOLDING For introduction to the TDI-based HR molded foam market, we targeted Active Polyol C that would allow foam producers to lower their added catalysts levels by 50%. We reasoned that this would significantly reduce worker exposure and lead to foam processes and products with much less volatile emissions. Improvement, or at least equivalence, in both foam processing and physical properties is important for the commercial success of any new product. Data demonstrating the performance of a molded foam grade of active polyol in these areas will be presented using the formulations given in Table 7. Note that for comparison purposes, we included a formulation based on the control polyols but at 50% reduced catalysts. At nominal packing levels, this formulation typically gives molded foams at a core density of 32 kg/m3 (2.0 lb/ft3).
Table 7. Formulations for low-density TDI HR molded foams. Component
SPECFLEX NC 630 Exp. Polyol C SPECFLEX NC 700 Niax Y-10184 Diethanol Amine, pure Dabco 33-LV Niax A-1 Water VORANATE T-80 Index Masterbatch Viscosity, mPa s at 25C Control At Full Catalysts 62.5 37.5 1.2 1.6 0.35 0.08 4.2 105 3450 Control At 50% Catalysts 62.5 37.5 1.2 1.6 0.17 0.04 4.2 105 3700 62.5 37.5 1.2 1.6 0.17 0.04 4.2 105 3600 Active Polyol C At 50% Catalysts

0.60 0.50 85 52 45.3 3.3 1.2 3.2 7.5 18.4

0.60 0.50 105 57 45.0 6.6 0.9 4.6 11.9 20.7

0.60 0.50 85 60 45.0 3.8 2.3 5.4 7.5 23.7

0.60 0.50 105 66 45.2 7.8 1.8 8.0 10.7 26

0.20 0.50 0.60 0.50 85 74 45.1 3.7 2.4 6.4 8.9 24.5

0.20 0.50 0.60 0.50 105 78 45.6 6.8 1.8 7.4 9.5 30.1

Reactivity
Gel time (s)

Properties
Molded density kg/m3 40 % CFD (Kpa) Airflow (cfm) 50 % CS (%) (CT) Humid aging 50 % HACS (%) CFD change 4th cycle (%) PVC staining

5.5

1.2

0.3

2.9

0.3

0.7

TEGOSTAB* is a trademark of Goldschmidt A. G.

All foam physical properties were measured according to VWs PV 3410 test methods. Catalyst A is N,N,NTrimethyl N-Hydroxyethyl-bis-Amino Ethyl Ether and Catalyst B is 2-Hydroxypropyl Imidazole. Dual-Hardness molded parts have been produced with active polyols and based on formulations similar to the one given in Table 6 without any difficulties. Square pads were also made with 100 parts of Polyol B and no catalysts with demolding times between 4 and 5 minutes, depending on the isocyanate index. Foam properties were found similar to the ones shown in Table 6 with Formulation C.

Also listed in Table 7 is the resultant masterbatch viscosity for each formulation. The masterbatch is that premixture of ingredients that will be combined with isocyanate in the mixing head in order to cause a foam producing reaction to begin. High viscosity here makes it difficult for the foam machine to accurately meter, mix and dispense the foam producing chemicals. A practical limit for many commercial foam machines is taken as 5000 mPa s. The formulation based on the new Active Polyol C is comparable to the two control formulations and well below the stated limit. Foam Processing The ability to benchmark important factors related to how well a foam formulation will process on foam making equipment is helpful in the development and commercial introduction of new products. With Dows in-house facilities and testing capabilities, candidate products undergo rigorous testing for the following list of processability factors.

Table 8. HR foam processability factors Masterbatch Viscosity Free-Rise Bulk Stability In-Mold Flowability Vent Stability Rise Profile Shrinkage Force-to-Crush Cure

viously established ratings. The data in Figure 4 show that all three foam formulations from Table 7 perform excellent in this test.
Free Rise Stability, Blair Units
Typical 9.5 9.5 Pass Fail 9.5

Foam Stability The production of defect-free molded foam parts requires that the candidate foam system possess an adequate level of stability through each stage of the foam reaction. The difficulty of maintaining adequate stability has increased as foam density has decreased and part designs have gotten thinner and more complex in shape. Free-Rise Bucket Foams An easy way to measure overall stability of the bulk foam is the free-rise bucket test illustrated in Figure 3.

10 9 8 7 6 5 4 3 2 1 0

Control At Full Catalysts

Control At 50% Catalysts

Polyol C At 50% Catalysts

Figure 4. Comparative free-rise stability performance.

Figure 5. In-mold flowability test. In this test, the mixture is dispensed into one end of a 38.1 x 76.2 x 6.3 cm (15 x 15 x 2.4 in) heated mold and purposely forced to flow over three rounded humps on its way to the vent holes at the opposite end of the mold. This stresses the growing foam and allows for comparison of candidate polyols on the basis of area percent voids on the show surface of the foam pad. Data for the current study are presented in Figure 6.
Flowability, Area % Voids
40
Fail Pass

Figure 3. Free-rise bucket test for bulk stability. In this test, a 22.7-liter (five imperial gallons) plastic bucket receives a shot from the foam-dispensing machine. The shot size is sufficient to totally fill the bucket with an approximate 10-centimeter (3.9-inch) crown of foam above the top of the bucket. Typical data recorded would include time to reach the top of the bucket, maximum and final foam height and an internal foam stability rating. The internal stability rating is obtained by cutting the cured foam in half (parallel to the direction of rise) and ranking the severity of voids on a 1 to 10 scale. A perfect foam would receive a 10 rating while a collapsed foam would be rated at 1. This grading is easily done by visual comparison of the candidate foam to a series of similar foams with pre-

10 9 8 7 6 5 4 3 2 1 0

Development Standard

Control At Full Catalysts

Control At 50% Catalysts

Polyol C At 50% Catalysts

Figure 6. Comparative flowability performance.

In this test, the control foam with the full level of added catalysts performs well below our internal development standard of no more than 10 area percent voiding. Dropping the catalyst levels of the control foam by 50% resulted

in the formation of voids over forty percent of the foams show surface. Active Polyol C at a 50% reduced catalysts loading performed better than the fully catalyzed control. Vent Stability Another common location for voids in a molded foam pad are those areas immediately below or adjacent to the vent devices. The data in Figure 7 were obtained by inspection and measurement of voids appearing below the vent holes in a simple 38.1 x 38.1 x 11.4 cm (15 x 15 x 4.5 in) heated test block mold. The vents for that mold were 1.6 millimeters (1/16 inch) in diameter and were arranged in a traditional five-hole pattern on the lid of the mold.
Vent Collapse, cm per vent 24 10 9 8 7 6 5 4 3 2 1 0 58
Rise Time, sec.

Foam Tightness The overall moldability of a foam is often said to be better with a tight formulation than a more open one. One consequence of running a tight formulation is the danger of foam shrinkage if the cell-walls are not adequately opened in a foam crushing event. The concern here is that the foam survive that crushing event without losing physical dimensions or bursting apart. If the foam is too tight, the internal stress forces during crushing may exceed the tensile strength of the fresh foam resulting in a physical splitting apart of the foam pad as it goes through the crushing device. This gross splitting results in a foam pad that must be scrapped. Comparison of relative shrinkage and force-tocrush performances are often useful in the development of new molded foam technologies. Shrinkage Our test scenario for measuring shrinkage in a standard test block of molded foam is illustrated in Figure 8.

26

Development Standard 7.2

Fail 4.2 Pass 2.3

Figure 8. Shrinkage testing.


Control At Full Catalysts Control

At 50%
Catalysts

Polyol C At 50%
Catalysts

Figure 7. Polyol C vent stability results.

The development standard we use is that no foam formulation should demonstrate more than 5 square centimeters of void area per vent. Above this level, the voids become easily detectable without cutting into the foam. For the present study, the control foam at full catalysis just meets this criterion. At the 50 percent reduced catalysts level, the control formulation is well into the unacceptable range of performance. Active Polyol C shows an improved level of performance. Also shown in this figure are the rise or mold fill times for each formulation. This number was recorded as the time for the first extrusion to be visible through the center vent hole. Twenty-four seconds is very typical for the fully catalyzed control formulation and that is matched nicely by the formulation containing Active Polyol C. The practical concern here is that any new formulation component not cause a faster rise time since most commercial foam plants have some limitations on how fast a mold lid can be closed and locked. Also, the rise time should not be so slow as for the other control formulation, so that the foam wont still be rising (and still be weak) as the molds go bumping along down the conveyor line.

In this test, foams are prepared in the normal manner, carefully removed from the mold and allowed to stand uncrushed overnight. The foam pad is then cut through the area of maximum vertical shrinkage, measured for minimum thickness and a percent shrinkage number calculated. Comparative data for the present work is presented in Figure 9.
Shrinkage, % 70 60 50 40
27 50
Fail Pass

Development Standard

30
20

20 10 0
Control At Full Catalysts Control At 50% Catalysts Polyol C At 50% Catalysts

Figure 9. Comparative shrinkage performance.

At the full catalyst levels, the control formulation performs at a shrinkage level that we consider to be an upper limit. With numbers above 50%, we have found that some commercial foamers have trouble running thicker foam pads through their crushing equipment. Shrinkage performance of experimental Active Polyol C is quite acceptable.

Force-To-Crush A useful laboratory test for comparing relative foam tightness is the so-called Force-to-Crush test. Our version of this test scenario is illustrated in Figure 10.

Cure For a new polyol to be commercially viable, it must contribute to an adequate level of foam cure at several critical points in the manufacturing process. Table 9 presents some practical examples of why foam cure is an important performance feature.
Table 9. Practical examples of foam cure.
Judged At Mold Fill Part Demold Station Crusher Station Shipping Station Affecting Bulk Stability and Flash Green Strength and Fingerprinting Hot Set Short Term Static Fatigue

The extent of cure at the moment of mold fill directly affects the amount of polymer that escapes the mold through the venting devices and parting line. If this is excessive, valuable material is wasted and additional pad trimming and cleanup labor may be needed. At the demold station, the foam must have enough green strength to resist tearing and fingerprinting due to human handling. These imperfections can usually be repaired offline at some additional expense. Another measure of cure comes after the foam leaves the crushing device. A poorly cured foam may densify and lose a few millimeters of its height dimension. Wrinkles may also appear on the exterior of the pad, further detracting from its cosmetic appearance. Similar events may occur at the shipping station when multiple pads are compressed into a shipping container for delivery to the seat assembly plants. A foam with inferior cure may exhibit a permanent loss in physical dimensions when removed from such shipping containers. The cure testing scenario used for this work is shown in Figure 12.

Figure 10. Force-to-crush test.

In this procedure, a foam is prepared in the normal manner, cured for the desired time and then removed from the mold (usually the standard test block mold) with care. The uncrushed foam pad is then quickly placed in the force measuring machine and deflected to 50% of its original thickness. Initial peak load and load decay are the basic data available. Figure 11 presents the initial peak load data for the present study. Foam based on the Active Polyol C falls nicely between the two control foams and all foams tested were well below any level of concern.
Force-To-Crush, N/323 cm 2 1,200 1,000 800 600 400
147 667 Development Standard 467 Fail Pass

Pour

Cure

Demold

Figure 12. The hot set cure test.

200 0
Control Control At 50%

Polyol C

At Full
Catalysts

At 50%
Catalysts

Catalysts

Figure 11. Comparative force-to-crush results.

In this test, foam pads made in the normal way are dispensed into a 38.1 x 76.2 x 6.3 cm (15 x 30 x 2.5 in) heated mold. After a 6 minute cure cycle, the foam pad was demolded and immediately roller crushed at a gap setting of 1.27 cm (1/2 in). The pad was then placed in the indenter jig where four indenter feet regionally deflect the foam to various residual thicknesses. Four replicate pads are tested at different deflection times and a final cumulative percent

hot set calculated. Data for the present study are presented in Figure 13.
Cure, % Hot Set 600 500 400
Development Fail Pass

tion values shown are taken from General Motors Engineering Standards Document GM6293M.
50% Cd Compression Set, % 50 40 30 65 93
Airflow, l/min Regular Set Humid Aged Set Typical Spec for Humid Aged Set Fail Pass

54

300 200 100 0 Control At Full Catalysts


145

267

240

20 10 0

15

17 13 10

Fail Pass

Typical Spec for Reg Set

16 12

Control At 50% Catalysts

Polyol C At 50% Catalysts

Control At Full Catalysts

Control At 50% Catalysts

Polyol D At 50% Catalysts

Figure 13. Comparative cure performance.

Figure 15. Compression set performance.

The control foam based on the high-functionality polyol, SPECFLEX NC 630 at a full catalyst level shows its typical outstanding cure performance. Lesser, but still very adequate, levels of cure were seen with reduced catalyst levels in both the control and Active Polyol C based foams. Foam Physical Properties All foam pads were thoroughly tested for physical properties according to G.M. test methods. Of particular interest are the load-bearing and compression set performances of the new Active Polyol C versus the control foams. Data comparing the 65% IFD (indentation force deflection) performance are presented in Figure 14.
65% IFD, N/323 cm 2 500
367

Historically, compression set specifications have been among the most difficult to meet. All three foam systems easily met the requirements of both the regular and humid aged specifications. Considering that compression sets usually respond to improvements in airflow, it seems logical that even better performance for the Active Polyol C based foams will be possible with formulation fine tuning. The remaining foam physical properties were consistent with those expected for a 32 kg/m3 (2.0 lb/ft3) core density molded foam based on TDI. Figure 16 presents a summary of foam physical properties wherein the performance of the Active Polyol C based foam is compared to that of the fully catalyzed control foam.
Foam Properties Loadbearing Regular CS Worse Same
5 20 6 7* 8

Better

400
327

342 Typical Performance

Humid Aged CS Tensile Tear Elongation Resiliency Airflow


17

300 200 100 0

Control At Full Catalysts

Control

Polyol C

* Still 150% Above specification

-25 -20 -15 -10 -5

10 15 20 25

At 50%
Catalysts

At 50%
Catalysts

Percent Change Relative To Control

Figure 14. Comparative loadbearing performance.

Figure 16. Performance features of active polyol versus fully catalyzed control foam.

Foams of this density grade and copolymer polyol level will generally give a 65% IFD in the range of 300 Newtons. Reducing the catalyst levels in the control foam gave the expected increase in loadbearing. Foams based on the new Active Polyol C were approximately 5% harder than the fully catalyzed control foam. Compression set data are presented in Figure 15. The figure shows the results for testing both the regular and the humid aged 50% Cd set. These foams fall within the General Motors Class C foam group and the typical specifica-

In this format, the data is intended to show, for example, that both compression sets are improved, as is the tear strength. The tensile strength of the experimental foam was 7% worse than the control but both foam systems were at least 150% over the minimum specification requirement for foams of this density. Elongation and resiliency performances were identical. Airflow in the experimental foam was 17% lower indicating a need for further formulation fine tuning. This formulation work is underway and current developments are confirming that total elimination of bis(2-Dimethylaminoethyl) ether, causing halovision (7), should be feasible.

6. CONCLUSION The present paper has demonstrated that amine emissions can be drastically reduced and even eliminated from flexible polyurethane foams through the use of new Dow active polyols combined or not with reactive catalysts. Further development of active polyols and adjustment of formulations are underway at Dow with the aim to give several options to the polyurethane industry to eliminate amine emissions. Volatile and odor tests, not reported here, are being carried out routinely to make sure these programs are giving the desired outcomes. The next step is the reduction of other volatile and fogging components in overall formulations in order to support the polyurethane industrys strive to very low emissions and to meet new stringent OEMs specifications. ACKNOWDLEGEMENT The authors wish to thank their many Dow colleagues who had a role in generating the data presented in this paper. Special thanks are extended to Alain Bleton, Lance Cooper, Richard Elwell, Olga Milovanovic, Chris Noakes, Olivia Renevey, Alan Schrock, Mark Sonnenschein, Antoine Storione, Ross Polk, Eve Taylor, Mary White, and the Meyrin and Freeport foam testing laboratories personnel. BIBLIOGRAPHY 1. Broos R.; Casati F.M.; Herrington R.M.(1999), Endurance of Polyurethane Automotive Seating Foams under Varying Temperature and Humidity Conditions, Journal of Cellular Plastics, Volume 36 May 2000, 207-245 Casati F.M.; Broos R.; Herrington R.M. ; Miyazaki Y. (1997), Tailoring the Performance of Molded Flexible Polyurethane Foams for Car Seats, Proceedings of the 1997 SPI Polyurethane World Congress, Technomics: Lancaster, PA., 402-420 Casati F.M.; Herrington R.M.; Broos R. ; Miyazaki Y.; Phan Thanh H and Cadolle D.; (1998) Automotive Seating Foams: Improving their Comfort and Durability Performance, Proceeding of Polyurethane the 1998 SPI Polyurethane conference; Technomics: Lancaster, PA., 417-431 Williams K.P.; J. Gerrard, D.L. The Degradation of Poly(Vinyl Chloride) Studied using Fourier-Transform Ramam Spectroscopy (1990), Eur. Polym. J., 26 (12)1355-8 Wilson A.S.; Gerrard D.L.; Bowley H.J. Problems of Interaction Between Plasticised PVC and Polyurethane Foam (1987). Presented at PVC 87, Brighton, UK., April 1987

7.

8.

9.

10.

11. 12.

13.

14.

2.

15.

3.

16.

4.

17.

5.

18.

19.

Christfreund A.; Huygens E.; Eling B. Amine-Free Catalyst Systems for Automotive Instrument Panels (1991), Proceedings of Polyurethane World Congress, Technomics: Lancaster, PA., 272-277 Polyurethane Amine Catalysts: Guidelines for Safe Handling and Disposal; Technical Bulletin AX173, November 2000; Alliance for the Polyurethane Indstry: Arlington, VA. Dernehl C.U. (1966), Health Hazards Associated with Polyurethane Foams, Journal of Occupational Medicine, 8/2, 59-62 Belin L.; Wass U.; Audunsson G. and Mathiasson L. (1983), Amines: Possible Causative Agents in the Development of Bronchial Hypperreactivity in Workers Manufacturing Polyurethanes from Polyisocyanates, British Journal of Industrial Medicine, 40, 251257 Bugler J.; Maddison P.; Hope S.J.; Mills I.G.; Nutt A.R. and Wright M.D. (1992) Tertiary Amine and Ethanolamine Catalysts in the Polyurethane Foam Industry: New Methods of Detection and Results of Personal Sampling in Nine UK Flexible Foam Factories, Cellular Polymers, 11/3, 171-200 Reed D. (1999), VW aims to cut amines, Urethanes Technology, 16/4, 3 Boinowitz T.; Burkhart G.; Schloens H.H. (1999) Emanations from Flexible Slabstock Foams Significant Reduction of Emissions is Now Possible, Proceeding of Polyurethanes Expo 99; Technomics: Lancaster, PA., 283-287 Zimmerman R.L.; Renken T.L. (1988): Low Odor Reactive Amine Catalysts for Polyurethane Foams, Proceedings of the SPI-31st Annual Technical/Marketing Conference, Technomics: Lancaster, PA., 36-40 Mercando L.A.; Kniss J.G.; Tobias J.D.; Plana A.; Listemann M.L.; Wendel S.; Non-Fugitive Catalysts for Flexible Polyurethane Foams, (1999) Proceeding of Polyurethanes Expo 99; Technomics: Lancaster, PA., 103-134 Kometani H; Tamano Y.; Masuda T.; Van Maris R.; Gay K.; (2000) The Investigation of Amine Emissions from Polyurethane Foam; Proceeding of Polyurethanes Conference 2000; Technomics: Lancaster, PA., 11-21 Rothe J.; Cordelair H.; Wehman C.; (2000) New Catalysts for Low VOC in Flexible Slabstock Foam, Proceeding of Polyurethanes Conference 2000; Technomics: Lancaster, PA., 109-116 Sikorski M.; Wehman C.; Cordelair H. (1999) New Additive Solutions for Low VOC in HR Molded Foams. Proceedings of Polyurethanes Expo 99; Technomics: Lancaster, PA., 135-144 Mapleston P; Riding Comfort is a Focus of Automotive Foam Systems (2001), Modern Plastics, March 2001, 30-31 Broos R.; Sonney J.M.; Phan Thanh H.; Casati F.M.; (2000) Polyurethane Foam Molding Technologies for 10

20.

21.

22.

23.

24.

Improving Total Passenger Compartment Comfort, Proceedings of the Polyurethane Conference 2000, Technomics: Lancaster, PA., 341-353 Malwitz N.; Manis A.; Wong S.W.; Frisch K.C.; (1986) Amine Catalysis of Polyurethane Foams, Proceedings of the 30th Annual Polyurethane Technical/Marketing conference, 1986, Technomics: Lancaster, PA., 338-353 Nutt A.R.; Skidmore D.W. (1987) Recovery of Chlorofluorocarbon 11 by Activated Carbon Scrubbing on a Polyurethane Foam Slabstock Plant, Cellular Polymers, Volume 6, 1987, 62-77 Herrington, R., De Genova, R., Casati, F. and M. Brown (1997), Molded Foams, Chapter 11, In Flexible Polyurethane Foams; Herrington, R. and K. Hock, eds., The Dow Chemical Company, Form No. 109-01061, 11.1-11.41 Herrington R.M. and Klarfeld D.R. (1983) Humid Aged Compression Set Phenomena in All Water Blown HR Molded Foams, Proceedings of the 1983 SPI Polyurethanes Conference, Technomics: Lancaster, PA., 177-182 Brasington R. and De Roeck H. (2000) Accelerated Testing for Durability Performance of Automotive Seating Foam, Proceedings of the API 2000 conference, Technomics: Lancaster, PA., 267-279

Jean-Marie Sonney
Dr. Jean-Marie Sonney graduated from the Swiss Federal Institute of Technology in Lausanne (Switzerland) and received his Ph.D. degree in physical organic chemistry from the same institution in 1979. After a postdoctoral research fellowship at the University of California, Santa Cruz, he joined the Geneva-based Research and Development group of BP Chemicals in 1981 and was transferred to the Dow Chemical Company in 1989. During these years, he had various responsibilities in the field of Analytical Chemistry, Quality Assurance, EH&S and Quality Management. He is currently a project engineer in the Development Group for Molded Foams and Specialties with responsibilities for NVH applications.

Johnson Tu
Johnson Tu is a Sr. Development Specialist in the Research and Development Group of The Dow Chemical Company in Freeport, Texas. He received a BS Science from the Chengkun University in Taiwan in 1983. He joined Dow in 1990 and has held several positions in R&D/TS&D and new Business Development.

BIOGRAPHIES Franois M. Casati Franois M. Casati graduated from ICPI (F), now CPELyon (Ecole Superieure de Chimie Physique Electronique de Lyon), in 1967. He has over 30 years of experience in Polyurethanes, Industrial Amines and Biocides, with S.N.P.E., Recticel, Abbott Laboratories, BP Chemicals and The Dow Chemical Company. During that time he has held various positions in Manufacturing, Marketing and Research & Development. He is currently a Product Development Leader, R&D, for the Flexible Foam business of Dow. Ronald M. Herrington
Ron Herrington is a Development Associate in the Polyurethane Products Technical Service and Development Laboratory of The Dow Chemical Company in Texas. He received a Bachelors Degree in Chemistry from the University of Houston in 1970. He joined Dow in that year and has since gained wide experience in the fields of slabstock and molded flexible foams.

Henri Mispreuve
Dr. Henri Mispreuve graduated from the State University of Mons (Belgium) in 1975 and received his Ph.D. in Physical Organic Chemistry from the same University in 1979. In 1984, he joined Dow Europe S.A. in Horgen (Switzerland). He is currently responsible for Technical Service and Development of Polyurethane Slabstock Foams at the companys International Development Center in Meyrin/Geneva (Switzerland).

Alain Fanget
Alain Fanget joined the Geneva-based Polyurethane Research and Development group of Union Carbide Europe in 1978 and was then transferred to BP Chemicals in 1979 and to The Dow Chemical Company in 1989 where he worked in various PU moulding and ACES applications. He is currently a Technical Service Specialist in the Automotive Thermosets R&D Group of the Meyrin International Development Center, with specific responsibilities in seating and NVH moulding applications.

11

For Additional Information:


The Dow Chemical Company Customer Information Group P.O. Box 1206 Midland, MI 48641-1206 Telephone (800) 441-4369 Website: www.dow.com

*Trademark of The Dow Chemical Company

Form No. 109-01552-901QRP

Вам также может понравиться