Sunteți pe pagina 1din 9

FULL PAPER

Amorphous Terpolymers www.mcp-journal.de

3D Printing Amorphous Polysiloxane Terpolymers


via Vat Photopolymerization
Justin M. Sirrine, Alisa Zlatanic, Viswanath Meenakshisundaram, Jamie M. Messman,
Christopher B. Williams, Petar R. Dvornic, and Timothy E. Long*

(PDMS) and −145  °C for polydiethylsilox-


Photocuring and vat photopolymerization (VP) additive manufacturing anes (PDESs).[1] They also exhibit excel-
(AM) is reported for two families of fully amorphous poly(dimethyl siloxane) lent chemical inertness, stability during
(PDMS) terpolymers containing either diphenylsiloxy (DiPhS) or diethylsiloxy ultraviolet (UV) radiation, low thermal
conductivity, high gas permeability, high
(DiEtS) repeating units. A thiol-functionalized PDMS crosslinker enables
thermal and thermo-oxidative stabilities,
rapid crosslinking in air using efficient thiol–ene addition. Differential low surface energies, and an unusually
scanning calorimetry and dynamic mechanical analysis (DMA) confirm the wide service temperature window.[1–5]
absence of crystallinity for the DiPhS-containing systems, while DMA shows Their extremely desirable low-temperature
a rubbery plateau extending to greater than 200 °C for the DiEtS-containing properties, in particular their pronounced
system. VP-AM of both photopolymer systems afford well-defined 3D geom- elasticity, arise from their inherent chain
flexibility derived from greater bond
etries, including high aspect ratio structures, which demonstrate feasibility angles and bond lengths of Si–O–Si (150°
of these photopolymers for the 3D printing of unique geometric objects that and 1.645 Å) units compared to C–O–C
require elastomeric performance to temperatures as low as −120 °C. units (113° and 1.412 Å), as well as from
very low intersegmental interactions.[1,2]
However, both PDMS and PDES remain
1. Introduction susceptible to low-temperature crystallization (at approx.
−40 to −50  °C and approx. −70  °C, respectively), which limits
Polysiloxanes remain distinct among other synthetic polymers their useful service temperature range to ≈10–20 °C above the
because of their unique combination of properties that result highest transition temperature for elastomeric applications.[6–11]
from inorganic –(Si–O)n– main chains and diverse organic Thus, crystallization suppression represents an extremely
pendant functionality. For example, polysiloxanes exhibit the attractive approach for improving low-temperature properties
lowest glass transition temperatures (Tg) known to polymer of polysiloxane elastomers (e.g., for aerospace applications),
science—for example, −123  °C for poly(dimethyl siloxane)s which extends their application temperature range to their
respective Tgs. As expected, the introduction of small amounts
of randomly placed, irregular repeating units sufficiently intro-
Dr. J. M. Sirrine, Prof. T. E. Long duces asymmetry and effectively reduces crystallizability.[6,12,13]
Department of Chemistry The literature provides many examples of this compositional
Macromolecules Innovation Institute effect.[6,14–16] Surprisingly, the addition of inorganic fillers (such
Virginia Tech, Blacksburg, VA 24061, USA
E-mail: telong@vt.edu
as silica, zinc oxide, iron oxide) is known to accelerate rather
V. Meenakshisundaram, Prof. C. B. Williams
than inhibit polymer crystallization, presumably due to fillers
Department of Mechanical Engineering acting as nucleating agents.[17]
Macromolecules Innovation Institute Recently, Zlatanic et al. reported the suppression of crystal-
Virginia Tech, Blacksburg, VA 24061, USA lization in PDMS through inclusion of only 3.6 mol% of sta-
A. Zlatanic tistically distributed diphenylsiloxy (DiPhS) repeating units,
Kansas Polymer Research Center or as low as 5 mol% of diethylsiloxy (DiEtS) units.[6] However,
Pittsburg State University
Pittsburg, KS 66762, USA they also reported that during silanoate-initiated ring-opening
Prof. P. R. Dvornic polymerization (ROP), polymers containing DiPhS units under-
Department of Chemistry went chain branching, producing phenyltrisiloxane branch
Pittsburg State University points in the polymer main chain with benzene serving as a
Pittsburg, KS 66762, USA leaving group. In contrast, polymers containing DiEtS repeating
Dr. J. M. Messman units did not show any branching under identical synthetic con-
Honeywell Federal Manufacturing & Technologies LLC
Kansas City, MO 64147, USA
ditions, making this system more suitable for low-temperature
elastomer applications due to a decreased likelihood of non-load
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/macp.201800425. bearing, dangling chains in the networks. In addition to dimeth-
ylvinylsilyl endgroups, both of DiPhS- and DiEtS-containing
DOI: 10.1002/macp.201800425 terpolymers also possessed ≈0.3 mol% of methylvinylsiloxy

Macromol. Chem.  Phys. 2019, 220, 1800425 1800425  (1 of 9) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mcp-journal.de

(MeViS) main-chain repeating units, providing a crosslinking thermal/thermomechanical properties. Most importantly,
functional greater than 2.0 for the resulting fully amorphous VP-AM enabled rapid construction of a variety of well-defined
polysiloxanes.[6] Furthermore, siloxane equilibration reactions 3D geometries, and optical imaging provided visual confirma-
occurred parallel to the ROP during synthesis and ensured ran- tion of the resolution of printed structures for various printed
domization of the repeating units. The same authors also sug- objects.
gested a mechanism for the polymerization of DiMeS–DiPhS-
containing polymers, which depicted (DiMeS)x and (DiPhS)y
blocky structures initially forming upon ROP of octamethylcy- 2. Experimental Section
clotetrasiloxane (D4) and octaphenylcyclotetrasiloxane (D4Ph2),
and subsequent redistribution culminating in polymer prod- 2.1. Materials
ucts with single DiPhS units separated by extended PDMS seg-
ments at long reaction times (≈200 min).[14] 2,2-Dimethoxy-2-phenylacetophenone (DMPA, 99%),
For the synthesis of elastomers, linear polysiloxane precur- diphenyl(2,4,6-trimethyl-benzoyl)phosphine oxide (TPO, 97%),
sors either undergo covalent crosslinking, most commonly and 2,5-bis(5-tert-butyl-benzoxazol-2-yl)thiophene (BBOT) were
using free radical initiators, hydrosilylation, or hydrolysis–con- purchased from Sigma-Aldrich and used as received. Poly­
densation reactions,[18–21] or reinforcement with fillers, in order (mercaptopropylmethylsiloxane-co-dimethylsiloxane) random
to improve mechanical properties.[2,22,23] Mechanical property copolymer (PDMS7.4k-SH) was purchased from Gelest (SMS-
improvements are also attainable with the synthesis of seg- 042) and used as received. PDMS7.4k-SH has a manufacturer-
mented, thermoplastic elastomers that undergo microphase reported Mn of 7000 g mol−1. The oligomer contained 5 mol%
separation, such as PDMS polyureas or PDMS-containing mercaptopropylmethylsiloxane repeating units, 95 mol%
polyimides.[3,22] Thiol–ene reactions also provide an efficient dimethylsiloxane repeating units, and a calculated value of 4.43
method for polysiloxane crosslinking. This “click” reaction pro- thiol-functional groups per chain based on Mn and composition.
ceeds with rapid rates and high yields, insensitivity to water
and oxygen (which both remain highly soluble in silicones),[21]
mild conditions, and without the formation of by-products.[24] 2.2. Synthesis and Characterization of PDMS7.0k-DiPhS
Furthermore, the step-growth polymerization kinetics of thiol– and PDMS7.5k-DiEtS
ene reactions allows for higher overall conversions prior to vit-
rification as compared to typical addition polymerization (e.g., All synthetic and characterization procedures for PDMS7.0k-
acrylates), which results in reduced polymerization shrinkage DiPh and PDMS7.5k-DiEt were provided earlier.[6,14] Charac-
and improved network homogeneity.[25] Consequently, the terization included online determination of refractive index
thiol–ene reaction offers attractive advantages for covalently increment (dn/dc), number-average molecular weight (Mn),
crosslinked polysiloxane elastomers.[26–32] weight-average molecular weight (Mw), dispersity (Đ), cyclic
The thiol–ene reaction has received wide attention in vat content (%), and Mark–Houwink parameters α and K, via gel
photopolymerization (VP) which is a class of layer-wise addi- permeation chromatography (GPC). Iodine (I2) titration meas-
tive manufacturing (AM) that produces solid objects from liquid ured vinyl content as grams I2 per 100 g polymer.
photopolymers upon irradiation with electromagnetic radiation
(e.g., UV light).[30,33–36] In this application, thiol–ene addition also
provides highly advantageous spatiotemporal control through 2.3. Determination of Polymer Vinyl and Thiol Content
formation of crosslinked networks only at the times and loca-
tions of direct UV exposure.[34,37] For example, a recent report Number of vinyl groups per polymer chain was calculated as
from Sirrine et al. described simultaneous chain extension and depicted in Equation (1).
crosslinking via thiol–ene coupling and acrylamide free radical
homopolymerization, which enabled a relatively low viscosity vinyl I2,g,exp M n ,GPC
# = * (1)
photopolymer to provide properties of higher molecular weight polymer chain I2,MW 100
precursors upon photocuring.[34] In another report, Wallin et al.
demonstrated VP of poly(mercaptopropylmethylsiloxane-co- This equation employs the iodine titration value in grams I2
dimethylsiloxane) and α,ω-divinyl PDMS, resulting in products per 100 g polymer (I2,g,exp), the molecular weight of I2 (I2,MW,
with strains at break over 400% and actuatable printed objects.[33] 253.81 g mol−1), a correction factor for the I2 titration normal-
However, the evaluation of sub-ambient thermal and mechanical izing it per 1 g polymer (100), and the polymer Mn as deter-
properties of the final products were not reported earlier.[33,34] mined by light scattering detection with GPC (Mn,GPC). This
To the best of our knowledge, the literature does not disclose provided calculated vinyl/chain values of 2.39 for PDMS7.0k-
any examples of 3D fabrication of ultra-low temperature elasto- DiPhS and 2.58 for PDMS7.5k-DiEt. The Gelest SMS-042
mers. Here, we demonstrate rapid (within 5 s) photocuring of product possessed trimethylsilyl endgroups that show identical
Zlatanic–Dvornic type, DiPhS- or DiEtS-containing, fully amor- 1
H NMR spectral resonances as backbone dimethylsiloxane res-
phous siloxane oligomers with a thiol-functional PDMS at low onances. Therefore, the number of thiol groups per chain was
photoinitiator content (e.g., 0.5 wt%). Thermogravimetric anal- calculated from manufacturer-provided theoretical molecular
ysis (TGA) probed thermal stability in nitrogen and in air, while weight (7000 g mol−1) and theoretical mol% thiol (5 mol%).
differential scanning calorimetry (DSC) and dynamic mechan- Two equations with two unknowns were formed, as shown in
ical analysis (DMA) assessed low-temperature transitions and Equations (2) and (3).

Macromol. Chem.  Phys. 2019, 220, 1800425 1800425  (2 of 9) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mcp-journal.de

DPSH and glass atop these three bottom layers, the sandwich was photo­
= 0.05 (2) cured under a UV lamp (Hanovia medium pressure Hg lamp, PC
DPSH + DPPDMS
451050; Ace Glass photochemical safety cabinet; 120 V, 60 Hz,
450 W UV power supply) for 3 min on each side. Samples were
7000 − 162.38 = DPSH * 134.27 + DPPDMS × 74.15 (3) used directly after film isolation from the photocuring sandwich.

These equations employed the degree of polymerization (DP)


of the mercaptopropylmethylsiloxane monomer repeating unit 2.5. Sample Preparation for Vat Photopolymerization
(DPSH) or the dimethylsiloxane repeating unit (DPPDMS), and
the molecular weights of the mercaptopropylmethylsiloxane Preparation of samples for VP was identical to photorhe-
repeating unit (134.27 g mol−1), dimethylsiloxane repeating ology studies except that TPO was used as a photoinitiator at
unit (74.15 g mol−1), total polymer molecular weight (7000 g 0.5 wt% due to spectral differences between the printer light
mol−1), and trimethylsilyl endgroups (162.38 g mol−1). source (BlueWave 75 UV spot curing lamp [Dymax 40078])
and the photorheology/photocalorimetry light source (Omni-
cure S2000). Both light sources contain broad-spectrum,
2.4. Sample Preparation for Photocuring and 3D Printing high-pressure mercury vapor bulbs, but the BlueWave 75 pos-
sesses slightly higher output at 405 nm than at 365 nm.[38] The
PDMS7.4k-SH (SH) and PDMS7.0k-DiPh (DiPhS) or PDMS7.5k- Omnicure S2000 output spectrum is available on the manufac-
DiEt (DiEtS) were added to 2-dram scintillation vials at stoichio- turer’s website and possesses roughly equal power at both 365
metric functional group concentration (e.g., 1:1 mol:mol based on and 405 nm. BBOT (0.1 wt%) was also added to the formula-
functional groups per chain), totaling 2.50 g, and all samples for tion to minimize cure-through during vat photopolymerization.
a single study were prepared at once. Separately, DMPA (1.00 g)
and chloroform (4.00 g) were added to a 2-dram scintillation vial
and homogenized with a vortexer for 20 s to form a photoiniti- 2.6. Analytical Methods
ator stock solution. Then, a microliter syringe was used to add
the proper amount of photoinitiator stock solution to the vortexed Proton nuclear magnetic resonance (1H NMR) spectroscopy
solutions such that DMPA loading occurred at 0.05, 0.1, 0.5, was performed using an Agilent U4-DD2 500 MHz NMR spec-
and 1.0 wt% for the SH/DiPhS and SH/DiEtS mixtures, totaling trometer and attached 96 sample robots. Photorheology was
eight samples. Finally, photoinitiated mixtures were mixed with conducted on a TA Instruments DHR-2 rheometer with Smart
a vortexer for 20 s until homogeneous and subsequently allowed Swap UV lower geometry, 20 mm quartz lower parallel plate,
to stand at room temperature, covered with aluminum foil, for 20 mm aluminum upper parallel plate, and an Omnicure S2000
2 h to ensure the absence of bubbles. Samples for TGA, DSC, high-pressure mercury light source with equipped 320–500 nm
and DMA were prepared by photocuring uniform films in a sand- filter. Samples were irradiated for 240 s at 8.5 mW cm−2, after
wich, structure depicted in Figure  1. Liquid photopolymer was measuring UV light output with a Silverline radiometer. Data
deposited inside an aluminum shim of defined thickness, placed was gathered at a 500 µm gap at 0.3% strain, 1 Hz in “Fast
atop a glass plate and layer of siliconized PET (silicone-coated Sampling” mode, enabling a 2 s−1 sampling frequency. UV irra-
Mylar). After placement of an additional layer of siliconized PET diation commenced after 30 s of dark oscillatory measurement.

Figure 1.  Photocuring assembly and procedure that provided uniform films of consistent thickness.

Macromol. Chem.  Phys. 2019, 220, 1800425 1800425  (3 of 9) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mcp-journal.de

The rheometer maintained an axial force of 0 N by making 3. Results and Discussion


slight adjustments in gap thickness, removing effects of polym-
erization shrinkage. Shear storage plateau moduli (GN0 ) were The literature provides many examples that describe the use of
determined as an average of the last 30 s shear storage moduli thiol–ene chemistry for photocuring of polysiloxanes. Müller
(G′) values. G′/G″ crossover time points, where G″ is loss mod- et al. described the photo-induced radical crosslinking of an
ulus, were determined using the “modulus crossover” function α,ω-ene-terminated polysiloxane with vinylsilyl endgroups and
in the accompanying TA Instruments TRIOS software. Soxhlet a thiol-containing polysiloxane, which showed nearly 100% thiol
extraction by THF was used to determine gel fractions of pho- and 98% ene functional group consumption and a first-order,
tocured discs produced with the photorheology method. Sox- linear relationship between the reaction rate (Rp) and the inci-
hlet samples were dried for 18 h under reduced pressure at 50 dent light intensity when stoichiometric mixtures were reacted
°C, weighed, extracted for 6 h by THF under reflux, dried for in air.[28] The system did not undergo ene homopolymerization,
18 h at 50 °C under reduced pressure, and weighed again. as indicated with copolymerization parameter products (r1 ∙ r2)
Photocalorimetry was conducted on a TA Instruments that were lower than 10−3, which is a value typical for copoly-
Q2000 differential scanning calorimeter (DSC) with the same merizations yielding alternating copolymer products.[28,40] The
light source as photorheology (Omnicure S2000) coupled with enthalpy of thiol addition across the olefin used was calculated
fiber-optic waveguide. To enable comparison to the 3D printing from Hess’s law and measured using DSC, providing values of
process, samples were weighed into TA Instruments Tzero 121 and 93 kJ mol−1, respectively.[28] Experimentally determined
sample pans at 20 ± 0.2 mg, resulting in crosslinked films with reaction enthalpy values were consistently lower than calculated
≈180 µm thickness. This closely mirrored the 3D printing pro- ones, likely due to incomplete conversion of ene groups. In
cess, where individual layers had a 150 µm thickness. Before another report, Cramer and Bowman described the carbon-
placing the sample into the cell, UV intensity was measured centered radical termination process in the presence of air,
with embedded sensors in the sample and reference cell posts. that is, oxygen, where growing polymer chains were oxidized
Once placed in the cell, samples were analyzed in modulated to peroxy radicals, which subsequently underwent hydrogen
DSC mode, equilibrated at 25 °C for 1 min, then irradiated at transfer to regenerate thiyl radicals.[41] Although these condi-
8.0 mW cm−2 for 6 min, and then held without UV irradiation tions likely reduce photoinitiator efficiency, Rp remained largely
for an additional 1 min to ensure proper return to baseline heat the same in air as well as in oxygen-free environments.[41]
flow. Polymerization exotherms were determined upon integra- Scheme 1 represents the UV-initiated curing of DiPhS- and/
tion of heat flow versus time curves. Standard DSC temperature or DiEtS-containing terpolysiloxanes used in this work with a
ramps were performed with a separate TA Instruments Q200 thiol-containing siloxane oligomer in the presence of a DMPA
DSC equipped with liquid nitrogen cooling system (LNCS) photoinitiator. Terpolymer compositions and characterization
which probed thermal transitions under constant heating/ information are summarized in Tables  1 and 2, respectively,
cooling rates of 10 °C min−1 and helium purge. TGA was per- and since trimethylsilyl endgroups of PDMS7.4k-SH could not
formed with a TA Instruments Q500 TGA at a heating rate of be distinguished in 1H NMR from dimethylsiloxy backbone
10 °C min−1 under constant N2 purge. DMA was performed resonances, the thiol content was accepted from manufacturer-
with a TA Instruments Q800 equipped with liquid nitrogen provided information. The reaction was evaluated in stoichio-
gas cooling accessory (GCA) and operated in oscillatory tension metric mixtures of DiPhS- and DiEtS-containing terpolymers
mode at 1 Hz, 0.1% strain, and 3 °C min−1 heating after equili- and SH crosslinker, by photocalorimetry and photorheology as
bration and 5 min isotherm at −150 °C. a function of photoinitiator concentration.
Figure 2A depicts polymerization exotherms from the photo-
calorimetry of DiPhS + SH system, plotted as heat flow versus
2.7. Vat Photopolymerization time. Integration of these curves yielded overall polymerization
exotherm, which remained nearly constant at 12–13 W g−1 as a
All VP information, including methods for generating the pho- function of photoinitiator concentration for all samples tested
topolymer working curve, is provided elsewhere.[34,39] (data not shown). Figure 2B displays time to reach heat flow

Scheme 1.  Crosslinking of amorphous 7.0k-DiPhS- or 7.5k-DiEtS-containing terpolysiloxanes with 7.4k-SH crosslinker and UV light in the presence of
photoinitiator.

Macromol. Chem.  Phys. 2019, 220, 1800425 1800425  (4 of 9) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mcp-journal.de

Table 1.  Compositions of amorphous siloxane terpolymers used. as a function of irradiation time from photorheology of the
DiPhS + SH system, reaching a plateau value (GN0 ) more rapidly
Sample Composition (theoretical) with increasing photoinitiator concentration. Finally, Figure 2D
DiMeS DiPhS DiEtS MeViS presents storage (G′) and loss (G″) modulus crossover, which
[mol%] [mol%] [mol%] [mol%] were shown earlier to indicate gel point for stoichiometrically
PDMS7.0k-DiPhS 96.1 3.6 0.0 0.3 balanced systems far above Tg,[42] as a function of photoinitiator
concentration. It was clear that G′/G″ reached a plateau value at
PDMS7.5k-DiEtS 94.7 0.0 5.0 0.3
higher DMPA concentrations, similar to the plateau in Rpmax.
Thus, in order to limit termination events and maximize UV pen-
maxima (s), corresponding to the maximum rate of polym- etration depth in the photopolymer, the lowest possible DMPA
erization (Rpmax), as a function of photoinitiator concentration, concentration that maximized Rpmax and minimized modulus
which reached a plateau value at the highest photoinitiator crossover G′/G″ time was chosen for VP-AM (e.g., 0.5 wt%).
concentrations (≈1 wt%), consistent with trends described in Thermal and thermomechanical properties of photocured
the literature.[39] Figure 2C shows shear storage modulus (G′) siloxane networks were evaluated on films made using the

Table 2.  Selected characterization data for siloxane terpolymers of Table 1.

Sample Mn, theor GPC MHS parameters Iodine value Vinyls per Dynamic
chain viscosity at 25 °C
dn/dc Mn Mw Đ Cyclic α K Theor Exp. Based on I2
(online) content exp and Mn,
GPC
[g mol−1] [mL g−1] [g mol−1] [g mol−1] – [%] – [mL g−1] grams I2 grams I2 per – Pa∙s
per 100 g 100 g
PDMS7.0k-DiPh 7069 −0.071 7042 12140 1.72 1 0.676 0.0148 8.0 8.6 2.39 0.21
PDMS7.5k-DiEt 6937 −0.086 7517 11110 1.48 2 0.692 0.0131 8.6 8.7 2.58 0.16

Figure 2.  A) Heat flow versus time from photocalorimetry at 25 °C, with UV light on at 0 min, as a function of photoinitiator (DMPA) concentration.
B) Time at peak heat flow from (A) versus photoinitiator concentration. C) G′ versus time from photorheology, with UV light on at 0 min. D) G′/G″
crossover time from photorheology as a function of photoinitiator concentration. Error bars represent standard deviation.

Macromol. Chem.  Phys. 2019, 220, 1800425 1800425  (5 of 9) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mcp-journal.de

sandwich geometry shown in Figure 1, which enabled pro- nitrogen than in air, and weight residues at the highest tem-
duction of free-standing films of uniform thickness. Sox- perature tested in air, was ≈20 wt%, consistent with the well-
hlet extraction of photocured films with THF indicated gel known thermal behavior of PDMS.[1,43] In nitrogen, the ulti-
fractions ranging from 85–88 wt% for all samples, but also mate weight residues were negligible, also consistent with the
resulted in film cracking, which precluded their thermome- available literature data.[1,43,44]
chanical analysis, a sensitive measure of polymer crystallinity. Figure  4A,B displays first heat DSC traces for photocured
As a consequence, unextracted films were used in the fol- films of DiPhS + SH and DiEtS + SH, respectively, together
lowing analyses. with data for the corresponding neat oligomers. While all
In nitrogen, TGA traces (see Figure  3A) showed higher values fall within the experimental error of the DSC method,
five-percent-weight-loss temperature (Td,5%) for photocured both photocured crosslinked networks displayed slightly higher
DiPhS + SH sample (Td,5%  = 408 °C) than for the photocured Tgs than their neat oligomers, consistent with the relatively
DiEtS + SH (Td,5%  = 390 °C), with both being higher than the low molecular weights of the later and the more pronounced
corresponding temperatures for neat DiPhS- and DiEtS-con- restrictions on the long-range segmental chain motions
taining terpolymers. In air, Figure 3B, both photocured DiPhS expected in covalently crosslinked networks from such precur-
+ SH and DiEtS + SH samples exhibited thermal stabilities sors. The photocured DiEtS + SH films displayed slightly lower
within the error of the instrument, showing Td,5% values of Tgs than films from the DiPhS + SH system, likely due to the
≈350 and 360 °C, respectively, also significantly higher than lower Tg of the neat DiEtS-containing terpolymer relative to the
the corresponding Td,5% values for neat oligomers. As expected, DiPhS-containing one (e.g., −128 versus −121 °C, respectively).
thermal stabilities of both photocured networks were higher in All samples demonstrated insensitivity to photoinitiator

Figure 3.  TGA of photocured DiPhS + SH and DiEtS + SH films overlaid Figure 4.  DSC first heating traces (10 °C min−1) for photocured films of
with neat DiPhS- and DiEtS-containing terpolymers and SH crosslinker (A) DiPhS + SH or (B) DiEtS + SH networks, overlaid with corresponding
in (A) nitrogen and (B) air. traces for neat oligomeric precursors.

Macromol. Chem.  Phys. 2019, 220, 1800425 1800425  (6 of 9) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mcp-journal.de

DiEtS + SH photocured film demonstrated a


significantly wider service temperature range,
extending its unchanged modulus up to 150 °C.
Figure  6A depicts a schematic of the layer-
wise VP-AM process employed for 3D printing
of these siloxane-containing terpolymer systems,
consisting of a UV light source, photopolymer
container, build stage attached to a linear actu-
ator, and dynamic mask, which enabled photo-
curing of entire layers at once. Figure 6B shows
the printing apparatus with various components
labeled, while Figure 6C,D shows various views
of a printed rook structure from the DiPhS +
SH siloxane-based photopolymer composition,
displaying a well-defined structure. Finally,
Figure 6E,F depicts the computer generated
STL file used to produce the object shown in
Figure 6C,D. The structure demonstrated the
combined ability of the printer and photopol-
ymer to support high aspect ratio structures and
limited cure-through. Although there were light
Figure 5.  Dynamic mechanical analyses (DMA) (3 °C min−1, 1 Hz) of photocured DiPhS + source and photoinitiator differences between
SH and DiEtS + SH films.
initial photocuring studies and VP (VP occurred
with the BlueWave 75 light source and TPO,
while the photocuring studies occurred with the
concentration, probed at the extremes of its values with photo- Omnicure S2000 and DMPA), no differences in mechanical
calorimetry and photorheology. Furthermore, samples did not properties between photocured and VP-produced objects were
display any evidence of crystallinity, or any thermal transitions expected. A previous study showed similar network Tg and
between Tg and 25 °C, where endo/exotherms characteristic for modulus for dental resins photocured with UV light intensi-
polysiloxane crystallization or melting would otherwise appear. ties varying over four orders of magnitude, and cure tempera-
Figure  5 shows thermomechanical response at constant tures varying by more than 60 °C.[45] VP occurred with 0.1 wt%
frequency (1 Hz) as a function of temperature for photocured BBOT, a UV absorber, which limited the cure-through that
DiPhS + SH and DiEtS + SH networks. Both samples under- occurred when printing this composition without UV absorber.
went an orders-of-magnitude drop in storage modulus (E′) as Aromatic rings of DiPhS repeating units may be responsible
a function of temperature, resulting in a pronounced peak for this phenomenon, since they will have different response to
in tan δ that corresponded to the primary α-relaxation, or Tg, UV light than the non-UV-active ethylsilyl units of the DiEtS +
in these polysiloxane networks. Photocured DiPhS + SH and SH photopolymer. We conclude that the avoidance of low-tem-
DiEtS + SH systems exhibited tan δ peaks at −113 and −118 °C, perature crystallization and UV absorbance, that is, a photocur-
respectively, and both were roughly 5 °C higher than the cor- able DiEtS + SH polymer system where the DiEtS terpolymer
responding Tg values determined by DSC. The DiPhS + SH is derived from a non-hydrolyzate source reagent (such as D3Et2
sample exhibited a consistent modulus from −90 to 80 °C, from or D4Et2 cyclic monomers described previously[6]) represents the
where it decreased until ≈150 °C. In contrast to this, the photo- most viable candidate for an ultra-low temperature elastomer
cured DiEtS + SH sample showed a small “bump” in modulus from 3D printing.
centered around −70 °C, which was attributed to crystallization
and melting of PDES segments expected to be present in this
DiEtS-containing terpolymer. This terpolymer was synthesized 4. Conclusions
from diethylsiloxane “hydrolysate,” a commercially available
mixture of various DiEtS cyclic oligomers and a moderately Two different amorphous dimethylsiloxane terpolymers, con-
high molecular weight PDES linear homopolymer, in a manner taining DiPhS and DiEtS repeating units, as well as MeViS
in which the siloxane equilibration did not fully randomize and chain-end vinylsilyl-functional groups, were crosslinked
the resulting polymer structure.[6] This resulted in retention of with a thiol-functional polydimethylsiloxane (SH). This
somewhat longer PDES segments in this terpolymer sample, yielded viable photopolymer systems for VP-AM via thiol–
which then exhibited the observed typical PDES behavior upon ene crosslinking chemistry when exposed to UV light in air.
thermomechanical analysis. As a consequence, it was expected Optimal photoinitiator concentration for VP-AM was deter-
that non-segmented, isolated DiEtS repeating unit containing mined by photocalorimetry and photorheology, and a pho-
siloxane terpolymers, which are obtained from pure cyclic tocuring procedure was developed to provide homogeneous
hexaethylcyclotrisiloxane (D3Et2) or octaethylcyclotetrasiloxane films of uniform thickness. TGA of films prepared from
(D4Et2),[6] would be the polymers of choice when complete crys- these polymer systems confirmed the expected high tem-
tallization inhibition is desired. It is also worth noting that the perature stabilities in both air and nitrogen, showing Td,5%

Macromol. Chem.  Phys. 2019, 220, 1800425 1800425  (7 of 9) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mcp-journal.de

Figure 6.  A) Schematic of the top-down scanning mask-projection vat photopolymerization (VP) apparatus employed for additive manufacturing
(AM) of siloxane terpolymers. B) VP-AM apparatus used in this work. C,D) 3D printed rook structures from DiPhS + SH photopolymer composition.
E,F) Computer-generated STL file used to create printed object shown in (C) and (D).

values ranging from 351 to 408 °C, while DSC showed com- DiPhS + SH system, as the phenyl-containing copolymers
plete absence of thermal transitions to 25 °C. DMA corrobo- are prone to uncontrolled branching,[6] which could affect
rated complete absence of crystallization or melting for the mechanical properties. Future work will instead include the
DiPhS + SH sample, with an orders-of-magnitude drop in use of DiEtS-containing terpolymers prepared from alternate
E′ at Tg. The photocured DiEtS + SH sample demonstrated sources of diethylsiloxane repeating units that provide com-
a small “bump” in E′, centered around −70 °C and attribut- plete crystallization suppression and no potential for chain
able to the crystallization and melting of longer than single branching, addition of UV blockers, and radical inhibitors
repeating unit PDES segments, expected to be present in (e.g., antioxidants) to provide vertical (z direction) and hori-
this sample as a side effect of the synthetic method used. zontal (x–y plane) resolution improvements, respectively,
Both photopolymer compositions demonstrated high ver- as well as printing in the presence of fillers, to improve the
satility through initial printing of various objects and high characteristically low mechanical properties of unfilled poly-
aspect ratio structures. Future work will avoid the use of the siloxane elastomers.

Macromol. Chem.  Phys. 2019, 220, 1800425 1800425  (8 of 9) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mcp-journal.de

Acknowledgements [17] M. F. Bukhina, S. K. Kurlyand, Low-Temperature Behaviour of Elasto-


mers, 1st ed., CRC Press, Boca Raton, FL 2007.
The authors graciously acknowledge funding from the Department of [18] S. C. Shit, P. Shah, Natl. Acad. Sci. Lett. 2013, 36, 355.
Energy’s (DOE) Kansas City National Security Campus, which is operated [19] P. Zheng, T. J. McCarthy, J. Am. Chem. Soc. 2012, 134, 2024.
and managed by Honeywell Federal Manufacturing & Technologies, LLC, [20] D. P. Melody, Br. Polym. J. 1989, 21, 175.
under Contract DE-NA-0002839. [21] F. de Buyl, Int. J. Adhes. Adhes. 2001, 21, 411.
[22] E. Yilgor, I. Yilgor, Polymer 2001, 42, 7953.
[23] I. Yilgor, T. Eynur, S. Bilgin, E. Yilgor, G. L. Wilkes, Polymer 2011, 52,
Conflict of Interest 266.
[24] C. E. Hoyle, C. N. Bowman, Angew. Chem., Int. Ed. 2010, 49, 1540.
The authors declare no conflict of interest. [25] M. Buback, H. Schroeder, H. Kattner, Macromolecules 2016, 49, 3193.
[26] L. M. Campos, T. T. Truong, D. E. Shim, M. D. Dimitriou, D. Shir,
I. Meinel, J. A. Gerbec, H. T. Hahn, J. A. Rogers, C. J. Hawker, Chem.
Mater. 2009, 21, 5319.
Keywords [27] C. F. Carlborg, T. Haraldsson, K. Oberg, M. Malkoch, W. van der
3D printing, elastomers, poly(dimethyl siloxane), siloxane terpolymers, Wijngaart, Lab Chip 2011, 11, 3136.
thiol–ene reaction, vat photopolymerization [28] U. Müller, A. Kunze, C. Herzig, J. Weis, J. Macromol. Sci., Part A
1996, 33, 439.
Received: September 22, 2018 [29] K. D. Q. Nguyen, W. V. Megone, D. Kong, J. E. Gautrot, Polym.
Revised: December 10, 2018 Chem. 2016, 7, 5281.
Published online: January 7, 2019 [30] L. M. Campos, I. Meinel, R. G. Guino, M. Schierhorn, N. Gupta,
G. D. Stucky, C. J. Hawker, Adv. Mater. 2008, 20, 3728.
[31] J.-S. Kim, S. Yang, H. Park, B.-S. Bae, Chem. Commun. 2011, 47, 6051.
[32] O. van den Berg, L.-T. T. Nguyen, R. F. A. Teixeira, F. Goethals,
[1] P. R. Dvornic, in Silicon-Containing Polymers: The Science and Tech- C. Özdilek, S. Berghmans, F. E. Du Prez, Macromolecules 2014, 47,
nology of Their Synthesis and Applications (Eds: R. G. Jones, W. Ando, 1292.
J. Chojnowski), Springer, Dordrecht, The Netherlands 2000, pp. [33] T. J. Wallin, J. H. Pikul, S. Bodkhe, B. N. Peele, B. C. Mac Murray,
185–212. D. Therriault, B. W. McEnerney, R. P. Dillon, E. P. Giannelis,
[2] W. Noll, Chemistry and Technology of Silicones, Elsevier, Cambridge, R. F. Shepherd, J. Mater. Chem. B 2017, 5, 6249.
MA 2012. [34] J. M. Sirrine, V. Meenakshisundaram, N. G. Moon, P. J. Scott,
[3] E. Yilgor, I. Yilgor, Prog. Polym. Sci. 2014, 39, 1165. R. J. Mondschein, T. F. Weiseman, C. B. Williams, T. E. Long,
[4] E. L. Warrick, O. R. Pierce, K. E. Polmanteer, J. C. Saam, Rubber Polymer 2018, 152, 25.
Chem. Technol. 1979, 52, 437. [35] I. A. Barker, M. P. Ablett, H. T. J. Gilbert, S. J. Leigh, J. A. Covington, J.
[5] F. Weinhold, R. West, Organometallics 2011, 30, 5815. A. Hoyland, S. M. Richardson, A. P. Dove, Biomater. Sci. 2014, 2, 472.
[6] A. Zlatanic, D. Radojcic, X. M. Wan, J. M. Messman, P. R. Dvornic, [36] S. C. Ligon, R. Liska, J. Stampfl, M. Gurr, R. Mülhaupt, Chem. Rev.
Macromolecules 2017, 50, 3532. 2017, 117, 10212.
[7] J. D. Helmer, K. E. Polmanteer, J. Appl. Polym. Sci. 1969, 13, [37] H. Leonards, S. Engelhardt, A. Hoffmann, L. Pongratz, S. Schriever,
2113. J. Bläsius, M. Wehner, A. Gillner, in SPIE Digital Library; Proc.
[8] S. J. Clarson, K. Dodgson, J. A. Semlyen, Polymer 1985, 26, 930. SPIE 9353, Laser 3D Manufacturing II 2015, 93530F, https://doi.
[9] A. Molenberg, S. Siffrin, M. Moller, S. Boileau, D. Teyssie, Mac- org/10.1117/12.2081169.
romol. Symp. 1996, 102, 199. [38] I. Marco-Rius, T. Cheng, A. P. Gaunt, S. Patel, F. Kreis, A. Capozzi,
[10] A. Molenberg, M. Moller, Macromolecules 1997, 30, 8332. A. J. Wright, K. M. Brindle, O. Ouari, A. Comment, J. Am. Chem.
[11] K. J. Miller, J. Grebowicz, J. P. Wesson, B. Wunderlich, Macromol- Soc. 2018, 140, 14455.
ecules 1990, 23, 849. [39] M. Hegde, V. Meenakshisundaram, N. Chartrain, S. Sekhar, D. Tafti,
[12] G. N. Babu, S. S. Christopher, R. A. Newmark, Macromolecules C. B. Williams, T. E. Long, Adv. Mater. 2017, 29, 1701240.
1987, 20, 2654. [40] G. Odian, Principles of Polymerization, 4th ed., Wiley, Hoboken, NJ
[13] K. A. Andrianov, G. L. Slonimskii, A. A. Zhdanov, Y. K. Godovskii, 2004.
V. Y. Levin, V. A. Moskalenko, J. Polym. Sci., Part A-1: Polym. Chem. [41] N. B. Cramer, C. N. Bowman, J. Polym. Sci., Part A: Polym. Chem.
1972, 10, 1. 2001, 39, 3311.
[14] A. Zlatanic, D. Radojcic, X. M. Wan, J. M. Messman, P. R. Dvornic, [42] H. H. Winter, Polym. Eng. Sci. 1987, 27, 1698.
Macromolecules 2018, 51, 895. [43] G. Camino, S. M. Lomakin, M. Lazzari, Polymer 2001, 42, 2395.
[15] L. Liu, S. Yang, Z. Zhang, Q. Wang, Z. Xie, J. Polym. Sci., Part A: [44] Z. C. Eckel, C. Zhou, J. H. Martin, A. J. Jacobsen, W. B. Carter,
Polym. Chem. 2003, 41, 2722. T. A. Schaedler, Science 2016, 351, 58.
[16] Y. Meng, J. Chu, J. Xue, C. Liu, Z. Wang, L. Zhang, RSC Adv. 2014, [45] L. G. Lovell, H. Lu, J. E. Elliott, J. W. Stansbury, C. N. Bowman,
4, 31249. Dent. Mater. 2001, 17, 504.

Macromol. Chem.  Phys. 2019, 220, 1800425 1800425  (9 of 9) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

S-ar putea să vă placă și