You are on page 1of 8

International Journal of Adhesion & Adhesives 34 (2012) 38–45

Contents lists available at SciVerse ScienceDirect

International Journal of Adhesion & Adhesives


journal homepage: www.elsevier.com/locate/ijadhadh

Comparison of a new, green foundry binder with conventional


foundry binders
John T. Fox a,n, Fred S. Cannon b, Nicole R. Brown c, He Huang a, James C. Furness d
a
Department of Civil and Environmental Engineering, Lehigh University, 13. E. Packer Avenue, Fritz Engineering Laboratory, Bethlehem, Pennsylvania 18015, United States
b
Department of Civil and Environmental Engineering, The Pennsylvania State University, 212 Sackett Building, University Park, Pennsylvania 16802, United States
c
Department of Forest Resources, The Pennsylvania State University, 202 Forest Resources Building, University Park, Pennsylvania 16802, United States
d
Furness-Newburge Inc., 376 Crossfield Drive, Versailles, KY 40383, United States

a r t i c l e i n f o a b s t r a c t

Article history: Metalcasting within the United States aims to meet ever-more stringent environmental standards as
Accepted 27 October 2011 new process technologies are developed. Conventional foundry core binders are responsible for up to
Available online 2 December 2011 70% of a foundry’s volatile organic compound (VOC) emissions. New core binder technologies are
Keywords: essential for environmental sustainability within foundries. Herein, conventional and novel foundry
High temperature adhesives core binders were appraised using thermal gravimetric analysis (TGA), dynamic mechanical analysis
Dynamic mechanical analysis (DMA), hot distortion testing (HDT), and (Pilot-scale) molten iron erosion tests. Inherently, these tests
Thermal analysis cannot replace full-scale casting trials to evaluate binder effectiveness, however, these tests were
Environmental issues performed to more fully elucidate binder properties that might cause casting defects or other unwanted
Foundry core binders
behaviors at high temperatures. During each of these lab protocols, the combination of collagen plus
alkali silicate as binders exhibited properties that matched or exceeded those of conventional phenolic
urethane. Also, in iron erosion testing, the collagen/alkali silicate binder exhibited the same low erosion
as conventional phenolic urethane. In hot distortion testing, the collagen–alkali silicate binder
exhibited longer resistance to thermal bending, and comparable thermal flexibility to conventional
phenolic urethane.
& 2012 Published by Elsevier Ltd.

1. Introduction poured into the mold. Upon metal solidification, the mold is
shaken-apart and the cast object is removed. Conventional foun-
Metalcasting is integral to American industry as castings are dry casting processes utilize green sand to create the outer
employed in nearly 90% of manufactured durable goods. Within geometry of the casting mold, and this is composed of a sand,
recent years, the metal casting industry has sought to meet ever- coal, water, and clay mixture. However, the sand, coal, and clay
more stringent environmental standards. However, conventional mixture is not suitable for creating molds for the more complex
core binders emit up to 70% of a foundry’s volatile organic metal casting that contain hollow cavities. The hollow cavities are
compound (VOC) emissions [1,2]. A number of foundry managers created using cores, as in the cylinders in an engine block. Cores
have expressed an interest in replacing these conventional core are composed of two components: sand and binder. Sand acts as a
binders with novel low-emission core binders—if they can offer filler, comprising 97–99% of a core. The sand is held together with
the same physical and mechanical properties at high tempera- 1–3% binder, or ‘core binder’ as is common foundry nomenclature.
tures. This paper describes the thermal and thermomechanical Herein, ‘binder’ will refer to the chemical mixture, which adheres
stability of novel, low VOC binders as an alternative to traditional the filler (i.e. sand) to form a core. Additionally, ‘core’ will refer to
binders. Inherently, these tests cannot supersede casting trials to the solidified ‘binder’ and ‘filler’ composite. These cores must
evaluate binder effectiveness, but they can certainly elucidate have adequate strength at room temperature and also maintain
nuances useful for understanding full-scale behavior. structural integrity when exposed to molten metal. During metal
Cores are important components of metalcasting molding pouring, the maximum temperatures experienced by cores range
operations. In the conventional casting process, a mold is shaped from 1510 1C at the molten iron interface, down to 50–300 1C
to form the geometry for a desired object, and molten metal is several inches away from this interface; temperature exponen-
tially declines from the metal interface through the core [3].
Paradoxically, after withstanding molten metal exposure and
n
Corresponding author. Tel.: þ1 610 758 2593. subsequent cooling, the binder must then disintegrate during
E-mail address: jtf211@lehigh.edu (J.T. Fox). shake-out, so that the sand can be easily removed from the cast

0143-7496/$ - see front matter & 2012 Published by Elsevier Ltd.


doi:10.1016/j.ijadhadh.2011.11.011
J.T. Fox et al. / International Journal of Adhesion & Adhesives 34 (2012) 38–45 39

product cavities. Subsequently, this shaken-out sand provides the collagen fibrils were reassembling. Over time, the amorphous
filler for the next cycle of green sand exterior molding. mineral phase transformed to nanocrystalline hydroxyapatite,
Several conventional petroleum-based binder systems, such as which then precipitated onto collagen fibrils. Yet, the utilization
phenolic urethanes and furans provide structural integrity, shake- of collagen and silicates for use in foundry applications is novel.
out propensity, and casting quality. However, these phenolic The authors hypothesized that the alkali–silicates will provide
urethanes and furan binders will thermally decompose after strength when exposed to molten metal, while the low-VOC
metal exposure, emitting as much as 30–70% of a foundry’s VOCs emitting collagen will gradually decompose after thermal expo-
and hazardous air pollutants (HAPs) [1,3]. For the paper herein, sure, as is conventionally desirable in the foundry industry, so as to
‘‘VOCs’’ are defined as the VOCs with C6–C16 hydrocarbon length, provide easy sand removal after metal solidification.
per foundry nomenclature [2]. It is these C6–C16 VOCs that may Alkali–silicates, specifically sodium silicates, are widely used
be considered a health concern. As expected, VOC and HAP as adhesives and are also commonly used as core binders in the
emissions increase proportionately with an increase in core- metalcasting industry [35–38]. Despite their many favorable
loading [4,1,5]. Recent work has determined that C6–C16 VOC characteristics, unmodified sodium silicate binders have one key
emissions for a collagen core binder are 70–90% less than for problem—difficulty in shakeout or the inability to breakdown
conventional phenolic-urethane binders [2,6]. Specifically, lab after metal solidification [38]. Other forms of alkali–silicates are
scale pyrolysis determined that collagen emits 41.5 mg VOC/g potassium and lithium silicates; most importantly lithium sili-
binder in the C6–C16 hydrocarbon range, compared to phenolic cates are the most insoluble following curing [36,37], which is
urethane’s 107.8 mg VOC/g binder in the C6–C16 range. Further, attributed to the smaller ionic radius of lithium than that of
collagen emits 7.7 mg HAP/g binder, compared to phenolic sodium or potassium [37]. Additionally, a blend of potassium and
urethane’s 32.1 mg HAP/g binder [2]. Instead of VOC and HAP lithium silicates provides a similar level of moisture resistance as
emissions, collagen releases more C1–C5 hydrocarbons than lithium silicate [37]. The increased moisture resistance provided
conventional phenolic urethane [2]. by lithium silicates is favorable for foundry core storage.
Despite the significant and necessary emission reductions The objectives of the work herein were to (1) develop a novel
enabled by collagen, it is only marginally employed in foundries. low-emission foundry core binder, (2) benchmark the relevant
The inability of collagen (alone) to withstand significant tem- physical as well as mechanical properties for both conventional
peratures imparted by molten iron (1290 1C–1510 1C) at the binders and novel binders that could influence casting quality, and
molten metal and sand interface is the key reason that collagen (3) provide a practical lab-scale test framework that can be utilized
is not implemented ubiquitously through the industry. Specifi- as a pre-cursor to full-scale molten metal tests that predict foundry
cally, this inability incurs erosion when molten iron pours onto binders’ ability to withstand molten metal erosion. The driving
the collagen (alone)-adhered core, and this produces costly hypothesis for this work is that collagen is a low-emission adhesive
casting defects. Collagen, which is available in large quantities that when used alone cannot provide the physical and mechanical
as a by-product of the meat-packing industry, is the fibrous properties necessary to withstand the high temperatures imparted
protein that strengthens skin, tendon and bones, and at ambient at the molten metal interface experienced by cores, however when
temperature, this triple-helical material has a tensile strength collagen is coupled and cross-linked with alkali–silicates the
equivalent to steel [7]. The individual collagen strands are 200– ‘hybrid’ binder can provide the necessary physical and mechanical
500 nm long, and each polypeptide strand is comprised of a three properties for use in metal casting.
amino acid repeating sequences. These typically include a repeat-
ing sequence similar to X–Y–Glycine, where Glycine comprises
about 34% of collagen, while the ‘‘X’’ and ‘‘Y’’ include proline 2. Materials and methods
(12%), alanine (10%), hydroxyproline (10%), glutamic acid (7%) or
other amino acids [8,9]. 2.1. Materials
When collagen is used as an adhesive, the triple helical structure
is denatured in water. Then as the adhesive dries, the polypeptide Sodium silicate was a modified sodium silicate provided by J.B.
strands randomly re-align and bond through Hydrogen Bonds [10]. DeVenne INC. (Berea, OH). This modified sodium silicate formula-
Hydrogen Bonds between the polypeptide strands are numerous, tion contained 42.5% by weight of dry solids. Collagen was
yet each bond is ‘weak’, with bond energy less than 25 kJ/mole and provided by Entelechy (Plymouth, MI), and came as a dry powder.
are at least an order of magnitude less than covalent bonds [11–13]. The phenolic urethane binder that was heat-set and cured by
The authors sought to increase the adhesive integrity at high triethylamine (TEA) was provided by Ashland Chemical Company
temperatures by increasing the bond energy [14] in a collagen (Columbus, OH), and came as a two-part binder that cured when
binder system by utilizing well-documented covalent intermolecu- mixed and heated. The blended potassium/lithium silicate was
lar organic cross-linking to increase the mechanical integrity and provided by PQ Corporation (Malvern, PA), and came as a liquid
stability of collagen [15–24]. Cross-linking also increases resistance with 29% solids content. Silica sand was provided by Wedron
to water solubility [25,26], which would improve core storage when Silica Company (Wedron, IL), and had a specified AFS grain
exposed to humidity. Of the types of cross-linking that could be fineness number ranging from 67 to 73.
employed, it has been the inorganic silicate cross-linking that we
found achieved more favorable results than organic cross-linking 2.2. Sample preparation overview
under the thermal duress imposed.
Specifically, the authors sought herein to include alkali–silicates Samples for thermogravimetric analysis (TGA), dynamic
in conjunction with collagen to produce a ‘hybrid’ organic/inor- mechanical analysis (DMA), and hot distortion testing (HDT) were
ganic binder. The crosslinking formation between collagen and prepared as described in respective test sections. In general, DMA
silicates has been studied by Coradin and others for biological and HDT samples were prepared by coating filler with binder, and
application [27–32]. This cross-link has the potential to increase the mix was molded into gum-stick sized core samples
the binding strength at both room temperature [33,27], and high (12.0070.30 mm  5.0070.15 mm  35.00 70.00 mm) for DMA,
temperature. Further work has been performed by Gelinsky et al. [34] or candy bar sized samples (15.24 cm  2.54 cm  1.27 cm) for
creating a mineralized collagen, which was made by precipitating HDT. The samples for DMA and HDT experiments contained the
amorphous calcium phosphate from a collagen solution as the typical percentages of sand (98–99%) and binder (1–2%) that
40 J.T. Fox et al. / International Journal of Adhesion & Adhesives 34 (2012) 38–45

would be used in a foundry setting. Core samples were prepared cured at 110 1C for 1 h, with the exception of heat cured phenolic
by mixing (1000 g) of washed foundry silica sand with the binder urethane, which was cured for 24 h.
system for 90 s at Level ‘2’ Solid State Speed Control, in a Hot distortion tests were performed with a custom hot
Commercial Kitchen Aid Mixer, Model Number KSMC50S (St. distortion test device. The samples were clamped in place with
Joseph, MI). 2.54 cm of sample clamped on an iron ledge, the Bunsen burner
TGA samples contained no filler (i.e. sand), since this allowed flame was centered on the exposed 12.7 cm of sample and the
us to track mass and heat changes more comprehensively, and micrometer monitored the deformation at a location that was
these samples were prepared with only the binder components as 1.27 cm from the end opposite the clamp. This deformation was
indicated in Table 1. monitored up to a maximum of 13.53 mm. The samples were
heated in the center of the test bar with a natural gas flame
2.2.1. Sample preparation for dynamic mechanical analysis (DMA) through 2.54 cm diameter perforated Bunsen burner, which had
For DMA analyses, the samples were hand packed into an an adiabatic flame temperature of 1960 1C at its source, and the
8-part wooden core mold. For the DMA test, the gum-stick sample perforated burner plate remained 2 cm from the bottom of the
size was 12.0070.30 mm  5.0070.15 mm  35.00 70.00 mm cantilevered sample. The micrometer transmitted deformation
(clamp determined the length). Samples were then cured at readings every second, and these readings were logged by the
110 1C for 1 h, with the exception of phenolic-urethane (that computer. For each parameter, 10–15 samples were analyzed via
was heat cured), which was cured for 24 h. The samples were HDT. The results exhibited inherent variability in both time to
then removed from the core mold and placed in a desiccator for given deflection, and deflection at failure (see discussion below).
overnight storage. The phenolic urethane samples that were TEA-
cured were provided to the Penn State team by Hitachi Metals 2.2.3. Sample preparation for thermal gravimetric analysis (TGA)
Automotive Components (HMAC) USA, LLC foundry, Lawrence- For TGA analyses, neat binder (no filler) samples were pre-
ville, PA. These samples were produced in core making machines pared, so that the thermal gravimetric profile for the binder
using 1.1% phenolic urethane, and were TEA cured in full-scale would not be masked by excessive sand mass that was thermally
operation. These production foundry cores had a 3 in. by 1 in. inert. For collagen samples, 5 g of collagen was mixed with water
diameter rod protruding from the main core. The Penn State team at a 1:3 mass ratio (collagen:water) at 70 1C. Once denatured, the
removed these rods, and manually filed them down to the collagen–water mixture was poured into an aluminum weighing
standard size used for DMA testing. pan to be dried at 110 1C for 12 h. For the modified sodium
For DMA analyses, samples were analyzed using a TA Instru- silicate, 10 g (4.3 g dry mass) was poured into an aluminum
ments DMA Q800 unit (New Castle, DE) with a dual cantilever weighing pan to be dried at 110 1C for 12 h. A hybrid binder
clamp. The samples were heated at 10 1C per minute to 500 1C, system with 10 g collagen and 8.7 g of K/Li silicate (dry mass
while the drive shaft oscillated with 5 mm amplitude at a basis) was prepared and dried at 110 1C for 12 h. For the other
frequency of 20 Hz. During testing, the samples were maintained hybrid binder system, 10 g of collagen was added into a solution
in a nitrogen atmosphere, to mimic the atmospheric conditions (8.5 g dry mass) of a modified sodium silicate, the entire mixture
experienced in the foundry core processing zone; the nitrogen was heated at 70 1C and stirred vigorously for 5 min. The hybrid
was purged at 2 L/min. binder mixture was then poured into an aluminum pan and dried
at 110 1C for 12 h. For the conventional phenolic urethane,
2.2.2. Hot distortion test (HDT) samples were prepared with 55 parts A to 45 parts B (by mass),
For hot distortion testing (HDT), the samples were formed into as is common industry practice; then dried at 110 1C for 24 h.
hand packed cores, in a 16-sample wooden core box, which For TGA tests, a small portion of dried binder (10–20 mg) was
created candy bar-sized cores that were 15.24 cm long, 2.54 cm removed with a scalpel to fit into the thimble sized alumina pan
wide and 1.27 cm thick. In this core box, the binder cores were used for TGA. The samples were placed in a TA instruments SDT

Table 1
Sample composition prepared for testing.

Name Composition for TGAn Composition for DMA and HDTn

Sodium silicate 4.3 g modified sodium silicate 12.8 g Na-silicate (dry mass), 1000 g sand
Collagen 10 g collagen, 30 mL water 10 g collagen, 1000 g sand
C þ K/Li silicate 10 g collagen, 8.7 g K/Li silicate (solids content) 10 g collagen, 8.7 g K/Li silicate (dry mass), 1000 g sand
C þ Na silicate 10 g collagen, 8.5 g Na-silicate (solids content) 10 g collagen, 8.5 g Na-silicate (dry mass), 1000 g sand
Phenolic urethane 0.55 parts A, 0.45 parts B 10 g phenolic urethane, 1000 g sand

n
All compositions on dry-mass basis.

Table 2
Molten iron erosion tests summary table for sand cores that were adhered as shown. (The balance of the core material was 98–99% silica sand).

Para-meter Sample 1% Collagen 1% PU 1.1% PU 1.3% Na Sil 0.85% Na Sil, 0.87% Li/K Sil, 0.29% Li/K Sil,
(Heat cured) (TEA-cured) 1% Col 1% Col 1% Col

Iron poured Kilogram 8.50 16.98 18.81 17.03 17.61 19.01 8.8
Hard-nessn Pre-pourn 77.5 73.5 85.5 76.0 78.5 85.0 80.5
Post-pourn 33.3 24.7 29.8 64.0 44.2 45.5 15.3
Erosion depth Total (cm) 3.08 0.36 0.33 oDetect o Detect o Detect 3.34
cm/kg Iron 0.362 0.021 0.018 oDetect o Detect o Detect 0.381
Sand erosion Total (gram) 44.19 12.30 7.92 oDetect o Detect o Detect 57.9
gram/kg Iron 5.20 0.72 0.42 oDetect o Detect o Detect 6.61
J.T. Fox et al. / International Journal of Adhesion & Adhesives 34 (2012) 38–45 41

2960—simultaneous DSC–TGA. The test commenced at room


Sodium Silicate
temperature in a 100% Argon atmosphere, with heating at 10 1C
per minute to 600 1C. Trials were tested in duplicate with the Phenolic Urethane
average between these duplicates representing the TGA curves Collagen
included herein. Duplicated results averaged a variation of Collagen + Sodium Silicate
slightly more than 5.8%. Collagen + Li/K Silicate
100
2.2.4. Sample preparation for pilot-scale molten iron erosion test
Cores were produced with (a) collagen, (b) sodium silicate,

Percent Mass Retained


(c) sodium silicate plus collagen, (d, e) potassium/lithium silicate 80
plus collagen (higher and lower silicate levels), (f) heat-cured
phenolic urethane (prepared at Penn State), and (g) tri-ethyla-
mine (TEA)-cured phenolic urethane (prepared at HMAC foundry). 60
Except for ‘‘g’’, all cores were prepared at Penn State; where they
were hand packed to be 12.7 cm by 12.7 cm wide and 5.1 cm
thick. These dimensions are with the exception of the TEA cured
40
phenolic urethane cores, which possessed a 7.6 cm by 7.0 cm flat
surface. This flat surface was extracted from a flat portion of a
production core, which originated from our supporting foundry,
HMAC. Cores were placed on a stand at a 451 angle. Molten iron, 20
which was at 2620–27201F (1438–1493 1C) was poured onto the 0 200 400 600 800
surface of the core from a height of 20.3 cm by passing the molten
Temperature (°C)
iron through a constant-head spout. The iron was melted in an
Inductotherm (Rancocas, NJ) Dura-Line electric induction furnace Fig. 1. Thermal gravimetric analysis for binders tested while heating at 10 1C per
in the Penn State pilot foundry; and 6.5–24.2. kilograms (14.3– minute to 500 1C while purged with argon, including; collagen, phenolic urethane
53.3 pounds) were poured onto the core. After running off the (heat-cured), sodium silicate, 1 part collagen and 0.85 parts sodium silicate, and
1 part collagen and 0.87 parts Li/K silicate.
angled core, the molten iron was captured in a ladle, then
weighed and poured back into the induction furnace. The cores
were then removed and evaluated for erosion and surface this combination also provided other favorable physical and
characterization. Molten iron erosion tests for conditions c, d, e, mechanical properties, as discussed below.
f, and g were tested in triplicate, while a, b, and e were tested in Upon request, we conducted tests with 1 part phenolic
duplicate. Results listed are the averages of these analyses. urethane with 0.87 parts sodium silicate. The TGA response for
The cores were evaluated for core scratch hardness before and this mix exhibited 25% mass loss at 420 1C, per Fig. 2. Also when
after the pour. In accordance with American Foundry Society forming cores these samples did not hold together very well, as
(AFS) 3318-00-S, scratch hardness test, which measures resis- described in Section 3.4.
tance to impression, by pressing the core surface with a Dietert
(Detroit, MI) Scratch Hardness Tester. The core scratch hardness 3.2. Dynamic mechanical analysis
was measured 2.5 cm from where the metal flowed off the core.
The erosion depth was measured using Mitutoyo calipers. Also, as DMA was performed to analyze the core’s physical properties at
a second measure of the erosion cavity that had been created high temperatures as well as to determine the glass transition
during iron pouring, displacement sand was poured into the temperature. The storage modulus and tan delta (tan d) figures are
erosion cavity and scraped level using a metal bar. Then this depicted herein. The storage modulus relates to the elastic compo-
displacement sand was poured into a pan and weighed to nent of material response; high modulus values indicate stiff
quantify the cavity of sand mass that had been eroded during materials, whereas lower values indicate compliant materials
metal pouring. [39–41]. In Fig. 3, we observe significant modulus drops, which
typically indicate the glass transition (Tg). Also, differences among
the collagen-based and non-collagen systems are apparent in the
3. Results and discussion 275–300 1C region; collagen and the collagen-hybrids reach mini-
mum modulus in this region, whereas phenolic urethane and
3.1. Thermal gravimetric analysis (TGA) sodium silicate show modulus plateaus. The plateaus suggest
crosslinked networks, and of the two, the sodium silicate is notably
Core binder thermal gravimetric analyses have exhibited the more stable than phenolic urethane. In fact, at approximately
mass retained by dry resins after exposure to high temperatures 350 1C, the modulus of the phenolic urethane samples became
(up to 800 1C). In Fig. 1, it can be seen that the inorganic sodium indistinguishable from those of either the neat collagen or the
silicate retained the most mass up to the highest temperatures collagen-hybrid systems. Note that all of the samples show slightly
(80% at 500 1C). Collagen retained the least mass (50% at 328 1C), increased modulus values from 350 1C to 500 1C, indicating the
while phenolic urethane (heat cured) retained 50% mass at 438 1C presence of a crosslinked network that has some thermal stability.
(Table 3). One of our early perceptions relative to shake-out Table 3 contains the percent storage modulus retained at
propensity was that it would be favorable to achieve a TGA mass 300 1C and 450 1C with respect to original storage modulus.
loss profile that roughly mimicked phenolic urethane, as such Sodium silicate retained 37.2% of the original storage modulus at
behavior might offer the more favorable simulated balance of 450 1C, while collagen retained only 4.8%. The collagen and alkali–
structural integrity during molten iron exposure and binder silicate hybrids retained 9.4 and 20.5%, for Li/K silicate and sodium
decomposition during shakeout. This balance of mass loss some- silicate, respectively. The phenolic urethane samples retained
what mimicked by collagen þsodium silicate, and collagen þLi/K 4.4 and 6.5% at 450 1C. The relative storage modulus retention
silicate (all three with 40% mass loss at 420 1C). As it turned out, indicates, which cores are most compliant at higher temperatures.
42 J.T. Fox et al. / International Journal of Adhesion & Adhesives 34 (2012) 38–45

Table 3
Summary Table providing key data from DMA, HDT, TGA, and Pilot-Scale molten iron erosion tests.

Test Parameter Binder material (along with 98–99% sand)

0.87% Li/K Sil, 0.85% Na Sil, 1% PU 1.1% PU 1.3% Na Sil 1% Coll-agen


1% Col 1% Col (Heat Cured) (TEA cured)

Molten iron Sand erosion g sand/kg iron poured nil nil 0.021 0.018 nil 0.362
test Erosion depth (cm/kg iron poured) nil nil 0.72 0.42 nil 5.2
Hardness, pre-pour 85 78.5 73.5 85.5 76 77.5
Hardness, post-pour 45.5 44.2 24.7 29.8 64 33.3
DMA Storage modulus at 300 1C (MPa) 48 72 189 53 362 22
Storage modulus at 450 1C (MPa) 179 164 70 46 683 99
T (1C) of start of deflection (start of loss of 200 200 125 70 145 225
stiffness)
Percent storage modulus retained at 300 1C (%) 2.4 7.9 12.0 7.5 19.9 1.0
Percent storage modulus retained at 400 1C (%) 9.4 20.5 4.4 6.5 37.2 4.8
Tan delta peak 0.51 0.42 0.38 0.31 0.63 0.75
Tan delta peak T (1C) 342 344 201 230 273 263
HDT Average time to 1 mm distortion (seconds)n 83 721 297 14 50 77 No test 227 5 65 710
Average time (s) to distortion of 13.5 mm (or 152 755 1977 137 85 7 24nnn No test 307 11 71 79nn
to break-failure)n

Samples with binder only


TGA T (1C) at 10% mass loss 273 267 244 No test 183 202
T (1C) at 50% Mass Loss 600 646 438 No test 183 328
% Mass retained at 750 1C 43 42 31 No test 84 24

n
Hot distortion test data lists average and standard deviation for 10–15 analyses.
nn
For collagen: 71 s to break-failure at 4 mm average distortion.
nnn
For phenolic urethane (hot cured), 3 samples broke before 13.5 mm distortion, while 10 samples did not even break at the maximum 13.5 mm distortion.

Sodium Silicate
Phenolic Urethane
Collagen
Collagen + Sodium Silicate
Collagen + L/K Silicate

120
100
2000
80
Storage Modulus (MPa)

60
40
1500
20
0
260 280 300 320 340 360
1000

500

0
100 200 300 400 500
Temperature (°C)
Fig. 2. Thermal gravimetric analysis for binders tested while heating at 10 1C per
minute to 500 1C while purged with argon, including; collagen, phenolic urethane Fig. 3. Dynamic mechanical analysis for core samples heated at 10 1C per minute
(heat-cured), sodium silicate, 1 part collagen and 0.85 parts sodium silicate, and to 500 1C while purged with 2 L/min of nitrogen, showing the storage modulus for
1 part phenolic urethane and 0.87 parts Na silicate. samples adhered with: (a) 1% collagen: (b) 1% phenolic urethane (heat-cured);
(c) 1.3% sodium silicate; (d) 0.85% sodium silicate plus 1% collagen; and (d) 0.87%
Li/K silicate plus 1% collagen. Insert: shows temperature regime from 250 1C to
The DMA tan d is a unitless parameter that corresponds with 375 1C (rubbery plateau), with storage modulus range from 0 to 125 MPa.
the tangent of the phase lag of material response. This phase lag
as defined as ratio of the loss modulus (E00 ; the component of the
modulus that has a time-dependent or viscous response) to the initial Tg. The secondary peak, occurring at higher temperatures,
storage modulus (E0 ; which represents the elastic or instanta- is typically broader, suggesting heterogeneity within the network
neous response). Peaks in the tan d represent damping events. and possibly an increased crosslink density. It is possible the
For polymers, this is typically interpreted as a relaxation, com- samples were incompletely cured, and that the Tg evolved during
monly a glass transition (Tg) [39–41]. the analysis. This would result in a second Tg peak at higher
Fig. 4 contains DMA tan d plots. Note that most samples temperature, which is more broad. An alternative possibility is
exhibit two peaks. The first peak generally correlates with the presence of distinct phases within the system, suggesting
drops in storage modulus (Fig. 3); hence, we interpret this as the inhomogeneity. Perhaps the material associated with the sand
J.T. Fox et al. / International Journal of Adhesion & Adhesives 34 (2012) 38–45 43

Sodium Silicate Sodium Silicate


Phenolic Urethane Phenolic Urethane
Collagen Collagen
Collagen + Sodium Silicate Collagen + Li/K Silicate
Collagen + Li/K Silicate Collagen + Sodium Silicate
14
0.8
12

0.6 10

Distortion (mm)
8
Tan Delta

0.4 6

0.2 2

0
0 50 100 150 200
0.0 Time (Seconds)
100 200 300 400 500
Temperature (°C) Fig. 5. Hot distortion test results for core samples adhered with: (a) 1% collagen;
(b) 1% phenolic urethane (heat-cured); (c) 1.3% sodium silicate; (d) 0.85% sodium
Fig. 4. Dynamic mechanical analysis for core samples heated at 10 1C per minute to silicate plus 1% collagen; and (e) 0.87% Li/K silicate plus 1% collagen. Cores were
500 1C while purged with 2 L/min of nitrogen, showing Tan delta for samples adhered heated over a natural gas flame.
with: (a) 1% collagen; (b) 1% phenolic urethane (heat-cured), (c) 1.3% sodium silicate;
(d) 0.85% sodium silicate plus 1% collagen; and (d) 0.87% Li/K silicate plus 1% collagen.

grains has unique intermolecular associations, which is reflected


in a second relaxation event. The sodium silicate sample only
shows one tan d, but it is unique in its breadth, suggesting a
highly heterogeneous network structure.

3.3. Hot distortion test

Core samples distorted as they were exposed to a natural gas


flame. Interestingly, we observed full distortion to the maximum
degree measureable with the instrument (13.5 mm) when we
appraised cores that were adhered with either phenolic urethane,
sodium silicate, collagen plus sodium silicate, or collagen plus
Li/K silicate. In contrast, collagen alone was the only core binder
that was unable to reach the maximum distortion (refer Fig. 5).
It is interesting to note that both sodium silicate and collagen
with lithium/potassium silicate blend were the cores that
remained undistorted for the longest time (more than 100 s
before 44 mm distortion). In a foundry setting, rapid distortion
at high temperatures provides unwanted core behavior; possibly
yielding casting defects. Yet further, the ability for the lithium/
potassium silicate core to withstand thermal stress the longest
until even distorting 1 mm is very favorable for many core Fig. 6. Erosion test block, adhered with 1 part collagen, 0.87 parts sodium silicate,
after experiencing 1482 1C molten iron.
scenarios in castings. The ability to withstand breakage until the
maximum measurable degree also corresponds with a cores
ability to withstand molten iron as demonstrated in Section 3.4. molten iron. The phenolic urethane cores eroded slightly, losing only
0.02 cm of core depth and 0.42 to 0.72 g of sand per kilogram of iron
3.4. Molten iron erosion test poured as in Table 2. Also, the hardness of the phenolic urethane
core decreased significantly when experiencing molten iron, to less
When the molten iron was poured onto cores, we observed that than half the sodium silicate core hardness. The collagen (alone)
the sodium silicate, and collagen plus silicate hybrid cores did not core exhibited the greatest erosion, even with the least amount of
erode (Fig. 6), whereas both phenolic urethane cores exhibited slight molten iron poured onto it. The collagen core erosion depth was
erosion, and the collagen (alone) cores exhibited significant erosion. approximately 20 times greater per kilogram of molten iron poured
The sodium silicate core decreased slightly in hardness after having than for phenolic urethane. However, it has been demonstrated in
17 kg of molten iron poured over the surface. The collagen plus these tests that when collagen is composited with either sodium
silicate hybrid cores exhibited more decrease in hardness than silicate or Li/K silicate, the molten metal erosion was prevented; and
did sodium silicate (alone) when experiencing a similar amount of the core performed as desired.
44 J.T. Fox et al. / International Journal of Adhesion & Adhesives 34 (2012) 38–45

sodium silicate distorted quickly during HDT, it was able to distort


to 13.5 mm without core failure. Also, a modified sodium silicate
bound core was able to experience molten iron pouring without
eroding or losing sand integrity. During TGA, heat cured phenolic
urethane retained 50% mass until 438.7 1C, while also retaining a
188.9 MPa storage modulus at 300 1C and also was able to distort to
13.5 mm without failure. This was similar to sodium silicate. Phenolic
urethane was also able to experience molten iron pouring, while only
eroding slightly. All tests indicated that collagen alone would perform
poorly as a core binder: it lost 50% mass by 328.3 1C during TGA
experienced the lowest storage modulus (19.76 MPa) value between
300 1C and 400 1C, collagen-bound cores only distorted 4.2 mm
before failure, and collagen eroded most significantly during molten
iron pouring. This can be attributed to the low bond energy associated
with hydrogen bonds. Conversely, the two novel collagen-silicate
binder systems compared favorably to conventional foundry binder
systems. During TGA, the novel binder systems retained 50% mass
until 600 1C or 646 1C, while also retaining a 19.0 or 50.6 MPa storage
modulus at 300 1C. The alkali silicate-collagen binders also possessed
Fig. 7. Hot distortion test bars, 1 part phenolic urethane with 0.87 parts sodium
a tan d peak similar to sodium silicate. These novel binders were able
silicate, exhibited extreme friability and were so fragile that they broke upon
removal from the mold. to distort 13.5 mm without failure during hot distortion tests. Most
importantly, during molten metal erosion tests, the two novel-silicate
binders experienced zero erosion.
The results herein demonstrate that collagen alone does not
provide sufficient physical and mechanical properties to work as a
core binder, which is attributed to the low bond energy attributed
to hydrogen bonds [10–14]. However, when collagen is used in
conjunction with lithium/potassium silicate or sodium silicate, a
foundry binder that possesses favorable physical and mechanical
properties is formed. Furthermore, the novel collagen plus silicate
binders compare favorably to conventional modified sodium sili-
cate and phenolic urethane. Yet, the novel collagen plus alkali–
silicate cores withstand thermal distortion more favorably than
phenolic urethane, which can help prevent unwanted core beha-
vior that contributes to casting defects. Overall, the novel collagen
and silicate binders possess: superior capabilities to withstand
distortion at high temperatures, appropriate mass retention at high
temperatures as well as most importantly withstand molten iron
erosion. Therefore, the results to date indicate that novel collagen-
silicate binders possess physical and mechanical properties at high
temperatures that are useful to foundries.
Fig. 8. Phenolic urethane and alkali silicate mixture, cyanoacrylate component of
phenolic urethane foundry binder rapidly polymerizes in the presence of the
water based alkali silicate mixture. Acknowledgments

As previously discussed, the authors were requested to pre- This research was supported in part by the National Science
pare cores using phenolic urethane and sodium silicate. However, Foundation (0824406, 0927967, and 0906271), Ben Franklin
mixing the oil-based phenolic urethane with the water-based Technology Partners and Hitachi Metals Automotive Components,
sodium silicate yielded a superficial blend of relatively disparate Lawrenceville, PA. The authors also acknowledge the very helpful
materials that exhibited little strength, as illustrated in Fig. 7. collaboration with Jim C. Furness and P. David Paulsen of Furness-
Further, it is believed that the cyanoacrylate component of the Newburge Co., and Jim Lamonski and Anhua Yu of HMAC foundry.
phenolic urethane polymerized when exposed to the water based Pilot scale molten iron erosion tests were poured in the Leonhard
sodium silicate caused the inability for the two binder systems to foundry pilot lab, with assistance by Robert C. Voigt, Matt
be utilized simultaneously (Fig. 8). Thus, the two systems are not Lumadue, Dan Supko and Randy Wells. 0824406, Standard Grant,
as compatible as the collagen and alkali silicate mixture. SGER/GOALI Industrial Personnel in Academe: Foundry Cores with
Reclaimed Sand and Novel Less-polluting Binders with UV Curing,
05/01/2008 to 04/30/2009. 0927967, Standard Grant, GOALI:
4. Conclusion Novel Low-Polluting Collagen–alkali Silicate, 08/15/2009 to 07/
31/2012. 0906271, Standard Grant, SGER/GOALI: Industry Pre-
The key physical and mechanical properties determined from the sence on Campus: Cross-linking of Collagen Binders for Foundry
work herein are summarized in Table 3. Sodium silicate performed Cores to Reduce VOC Emissions, 10/01/2009 to 12/31/2010.
very well during TGA as it retained 78.7% mass to 800 1C, yet this may
contribute to the poor shake-out characteristics. DMA tests indicated References
that sodium silicate had the greatest storage modulus throughout the
entire test, specifically having a 362.4 MPa storage modulus at 300 1C, [1] Glowacki CR, Crandell GR, et al. Emissions studies at a test foundry using an
when collagen only possessed a 21.64 MPa modulus. Although advanced oxidation clear water system. Am Foundry Soc Trans 2004:3.
J.T. Fox et al. / International Journal of Adhesion & Adhesives 34 (2012) 38–45 45

[2] Wang Y, Cannon FS, et al. Characterization of hydrocarbon emissions from [22] Heijmen FH, du Pont JS, Middelkoop E, Kreis RW. Cross-linking of dermal
green sand foundry core binders by analytical pyrolysis. Environ Sci Technol sheep collagen with tannic acid. Biomaterials 1997;18:749–54.
2007;41(22):7922–7. [23] Taguchi T, Saito H, et al. Bonding of soft tissues using a novel tissue adhesive
[3] Wang Y, Cannon FS, Voigt RC, Komarneni S, Furness JC. Effects of advanced consisting of a citric acid derivative and collagen. Mater Sci Eng: C
oxidation on green sand properties via iron casting into green sand molds. 2004;24(6-8):775–80.
Environ Sci Technol 2006;40:3095–101. [24] Kanth VS, Ramaraj A, Rao JR, Nair BU. Stabilization of type I collagen using
[4] Goudzwaard JE, Kurtti CM, et al. Foundry emissions effects with an advanced dialdehyde cellulose. Process Biochem 2009;44:869–74.
oxidation blackwater system. Am Foundry Soc Trans 2004;79(3):20. [25] Gupta B, Hilborn J, Bisson I, Frey P, Plummer C. Thermal crosslinking of
[5] Schifo JF, Radia JT, Crandell GR, Mosher G. Iron foundry hazardous air collagen immobilized on PAA grafted poly(ethylene terephthalate) films.
pollutants, what we know and what we don’t. Am Foundry Soc Trans J Appl Polym Sci 85, 1874;2002.
2003;111. [26] Pizzi A, Mittal KL. Handbook of Adhesive Technology. 2nd ed. New York:
[6] CERP CERP-Technicon, LLC. Comparison of Binder Systems relative to Emis- Marcel Dekker Inc.; 2003.
sions (Orally presented data; on CERP web page); (2000). [27] Coradin T, Durupthy O, et al. Interactions of amino-containing peptides with
[7] Buehler MJ. Nature designs tough collagen: explaining the nanostructure of sodium silicate and colloidal silica: a biomimetic approach of silicification.
collagen fibrils. Proc Natl Acad Sci U S 2006;103(33):12285–90. Langmuir 2002;18(6):2331–6.
[8] Palfi VK, Perczel A. How stabile is a collagen triple helix? an ab initio study on [28] Allouche J, Boissiere M, Helary C, Livage J, Coradin T. Biomimetic core-shell
various collagen and b-sheet forming sequences J Comput Chem 2007;29(9). gelatine/silica nanoparticles: a new example of biopolymer-based nanocom-
[9] Eastoe JE, Leach AA. In: Recent advances in gelatin and glue research. New posites. J Mater Chem 2006;v 16(30):3120–5.
York: Pergamon Press;1958. [29] Eglin D, Mosser G, Giraud-Guille M-M, Livage J, Coradin T. Type I collagen, a
[10] Stainsby G. Recent advances in gelatin and glue research. New York: versatile liquid crystal biological template for silica structuration from nano
Pergamon Press; 1958. to microscopic scales. Soft Matter 2005;1(2):129–31.
[11] Emsley J. Very strong hydrogen bonding. Chem Soc Rev 1980;9:91–124. [30] Eglin D, Shafran KL, Livage J, Coradin T, Perry CC. Comparative study of the
[12] Buehler MJ, Ackbarow T. Fracture mechanics of protein materials. Mater influence of several silica precursors on collagen self-assembly and of
collagen on ‘Si’ speciation and condensation. J Mater Chem 2006;16(43):
Today 2007;10(9):46–58.
4220–30.
[13] IUPAC. Hydrogen bond, compendium of chemical terminology, 2nd ed. (the
[31] Coradin T, Livage J. Synthesis, characterization and diffusion properties of
Gold Book). Compiled by McNaught AD, Wilkinson A. Oxford: Blackwell
biomimetic silica-coated gelatine beads. Mater Sci Eng C 2005;v 25(2):201–5.
Scientific Publications; 1997. XML on-line corrected version: /http://gold
NATO Advanced Study Institute. ASI on Learning from Nature How to Design
book.iupac.orgS; 2006. Created by Nic M, Jirat J, Kosata B. Updates compiled
New Implantable Biomaterials: From Biomineralization Fundamentals..
by Jenkins A. doi:10.1351/goldbook.
[32] Coradin T, Marchal A, Abdoul-Aribi N, Livage J. Gelatine thin films as
[14] Strong AB. Fundamentals of Composites Manufacturing: Materials, Methods,
biomimetic surfaces for silica particles formation. Colloids Surf B (Biointer-
and Applications. Dearbron, Michigan: Society of Manufacturing Engineer-
faces) 2005;44(4):191–6.
ings; 2008.
[33] Gonzalez L, Rodriguez A, Ibarra L. Reinforcing effect of a magnesium-silicate
[15] Chen Y, Mak AFT, et al. Formation of apatite on poly(alpha-hydroxy acid) in
treated with a 4-amino methyl pyridine in Epichlorhydrin rubber (Eco)
an accelerated biomimetic process. J Biomed Mater Res Part B-Appl Biomater vulcanizates. J Appl Polym Sci 1991;43(7):1327–33.
2005;73B(1):68–76. [34] Gelinsky M, Welzel PB, et al. Porous three-dimensional scaffolds made of
[16] Charulatha V, Rajaram A. Influence of different crosslinking treatments on mineralised collagen: preparation and properties of a biomimetic nanocom-
the physical properties of collagen membranes. Biomaterials 2003;24: posite material for tissue engineering of bone. Chem Eng J 2008;137(1):
759–67. 84–96.
[17] Usha R, Ramasami T. Effect of crosslinking agents (basic chromium sulfate [35] Rabbii A. Sodium Silicate Glass as an Inorganic Binder in Foundry Industry.
and formaldehyde) on the thermal and thermomechanical stability of rat tail Iran Polym J 2001;10(4).
tendon collagen fibre. Thermochim Acta 2000;356:59–66. [36] Lambourne R, Strivens TA. Paint and surface coatings, theory and practice
[18] Wollensak G, Spoerl E. Collagen crosslinking of human and porcine sclera. .2nd ed. Cambridge, England: Woodhead Publishing Ltd.; 1999 pp. 393, 394.
J Cataract Refract Surg 2004;30(3):689–95. [37] Veinot DE, Langile KB, Nguyen DT, Bernt JO. Fire protective coatings based on
[19] Verzijl N, DeGroot J, et al. Crosslinking by advanced glycation end products the Na/K/Li silicate system. J Can Ceram Soc 1990;59(4):32–6.
increases the stiffness of the collagen network in human articular cartilage—a [38] Srinagesh K. Chemistry of sodium silicate as 2 sand binder. Int Cast Metals J
possible mechanism through which age is a risk factor for osteoarthritis. 1979;4(1):50–63.
Arthritis Rheum 2002;46(1):114–23. [39] Sepe MP. Dynamic mechanical analysis for plastics engineering. Norwich,
[20] Kato YP, Silver FH. Properties of manually produced and automated contin- NY: Plast Des Libr; 1998 pp 1–35.
uous collagen fibers. IEEE Eng Med Biol Soc 11th Ann Int Conf 1989. [40] PerkinElmer, Inc. Introduction to dynamic mechanical analysis (DMA).
[21] Harriger MD, Supp AP, et al. Glutaraldehyde crosslinking of collagen PerkinElmer, Inc. Waltham MA; 2008.
substrates inhibits degradation in skin substitutes grafted to athymic mice. [41] Menard KP. Dynamic mechanical analysis, a practical introduction. 2nd ed.
J. Biomed Mater Res 1997;35(2):137–45. Boca Rotan, FL: CRC Press; 2008.

You might also like