Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


We rationalize the unusual gas transport behavior of polymer-grafted nanoparticle (GNP) membranes. While gas permeabilities depend specifically on the chemistry of the polymers considered, we focus here on permeabilities relative to the corresponding pure polymer which show interesting, "universal" behavior. For a given NP radius, Rc, and for large enough areal grafting densities, σ, to be in the dense brush regime we find that gas permeability enhancements display a maximum as a function of the graft chain molecular weight, Mn. Based on a recently proposed theory for the structure of a spherical brush in a melt of GNPs, we conjecture that this peak permeability occurs when the densely grafted polymer brush has the highest, packing-induced extension free energy per chain. The corresponding brush thickness is predicted to be h max = 3 R c , independent of chain chemistry and σ, i.e., at an apparently universal value of the NP volume fraction (or loading), ϕNP, ϕNP,max = [Rc/(Rc + hmax)]3 ≈ 0.049. Motivated by this conclusion, we measured CO-2 and CH4 permeability enhancements across a variety of Rc, Mn and σ, and find that they behave in a similar manner when considered as a function of ϕNP, with a peak in the near vicinity of the predicted ϕNP,max. Thus, the chain length dependent extension free energy appears to be the critical variable in determining the gas permeability for these hybrid materials. The emerging picture is that these curved polymer brushes, at high enough σ behave akin to a two-layer transport medium - the region in the near vicinity of the NP surface is comprised of extended polymer chains which speed-up gas transport relative to the unperturbed melt. The chain extension free energy increases with increasing chain length, up to a maximum, and apparently leads to an increasing gas permeability. For long enough grafts, there is an outer region of chain segments that is akin to an unperturbed melt with slow gas transport. The permeability maximum and decreasing permeability with increasing chain length then follow naturally.

Free full text 


Logo of nihpaAbout Author manuscriptsSubmit a manuscriptHHS Public Access; Author Manuscript; Accepted for publication in peer reviewed journal;
Macromolecules. Author manuscript; available in PMC 2023 Apr 26.
Published in final edited form as:
Macromolecules. 2022 Apr 26; 55(8): 3011–3019.
Published online 2022 Apr 6. https://doi.org/10.1021/acs.macromol.1c02367
PMCID: PMC9377655
NIHMSID: NIHMS1829350
PMID: 35978703

Understanding Gas Transport in Polymer-Grafted Nanoparticle Assemblies

Abstract

We rationalize the unusual gas transport behavior of polymer-grafted nanoparticle (GNP) membranes. While gas permeabilities depend specifically on the chemistry of the polymers considered, we focus here on permeabilities relative to the corresponding pure polymer which show interesting, “universal” behavior. For a given NP radius, Rc, and for large enough areal grafting densities, σ, to be in the dense brush regime we find that gas permeability enhancements display a maximum as a function of the graft chain molecular weight, Mn. Based on a recently proposed theory for the structure of a spherical brush in a melt of GNPs, we conjecture that this peak permeability occurs when the densely grafted polymer brush has the highest, packing-induced extension free energy per chain. The corresponding brush thickness is predicted to be hmax=3Rc, independent of chain chemistry and σ, i.e., at an apparently universal value of the NP volume fraction (or loading), ϕNP, ϕNP,max = [Rc/(Rc + hmax)]3 ≈ 0.049. Motivated by this conclusion, we measured CO-2 and CH4 permeability enhancements across a variety of Rc, Mn and σ, and find that they behave in a similar manner when considered as a function of ϕNP, with a peak in the near vicinity of the predicted ϕNP,max. Thus, the chain length dependent extension free energy appears to be the critical variable in determining the gas permeability for these hybrid materials. The emerging picture is that these curved polymer brushes, at high enough σ behave akin to a two-layer transport medium – the region in the near vicinity of the NP surface is comprised of extended polymer chains which speed-up gas transport relative to the unperturbed melt. The chain extension free energy increases with increasing chain length, up to a maximum, and apparently leads to an increasing gas permeability. For long enough grafts, there is an outer region of chain segments that is akin to an unperturbed melt with slow gas transport. The permeability maximum and decreasing permeability with increasing chain length then follow naturally.

Polymeric membranes have been extensively studied to effectively and selectively separate gas mixtures in a variety of applications such as O2 from air and in natural gas purification.1-7 A membrane’s separation performance is characterized by its gas permeability (Pi), and its ideal selectivity of gas i relative to gas j (αi,j = Pi/Pj). In the framework of the solution diffusion model, the permeability is defined as the product of the penetrant’s diffusion (Di) and solubility coefficients (Si),

Pi=Di×Si;αi,jPiPj=DiDj×SiSj
(1)

An optimal membrane for a given gas pair, e.g., CO2 and CH4, combines high CO2 flux (permeability) with high CO2/CH4 selectivity, i.e., CO2 purity in the permeate. However, typical membranes exhibit a trade-off between permeability and selectivity, captured by the empirical Robeson Upper Bound, e.g., Figure 1.8-11 The grey points in this figure are different pure polymer membranes tested for CO2/CH4 separations. Such “Robeson” plots have been extensively used to compare the performance of different membrane materials with the goal of finding polymers which exceed those with best current performance. Many polymers with varying chemistries have been developed, but only a few, such as thermally rearranged (TR) polymers12 and polymers of intrinsic microporosity (PIMs),13,14 show the desired, significant increases in Pi and αi,j. Since synthesizing and testing these newer classes of polymers is an expensive and time-consuming process, it is critical to explore a variety of other constructs that could offer performance advantages. Here, we focus on one such framework, i.e., when polymers are mixed with inorganic particles to form “mixed-matrix membranes”. These hybrid materials have increased mechanical strength,15-19 ease of processing, and improved temporal stability20-22 relative to the neat polymer — all desirable traits. However, incorporating particles into a polymer generally has deleterious consequences on gas transport. Conventional composite theory, e.g., Maxwell’s theory, predicts that a polymer matrix mixed with impermeable spherical filler (with volume fraction ϕNP) will have reduced permeability relative to the pure polymer:23,24

PϕPn=1ϕNP1+ϕNP2
(2)

An external file that holds a picture, illustration, etc.
Object name is nihms-1829350-f0001.jpg

Robeson plot illustrating the trade-off between CO2/CH4 selectivity and CO2 permeability. Open grey symbols are literature data for a variety of polymers studied over the last 30 years. The open black square represents pure poly(methyl acrylate) [PMA], while colored symbols are PMA-grafted NPs at constant areal graft density, σ ≈ 0.47 chains/nm2 but different graft molecular weight Mn. The filled black circles are physical mixtures of PMA and silica NPs. The “upper bound” is an empirical description of the best available membranes as of 2008.

Pϕ and Pn are the permeabilities of the mixed-matrix and pure polymer, respectively. The numerator reflects the decrease in gas solubility due to the presence of the filler, while the denominator represents the increased tortuosity encountered by the gas in the filler/polymer mixture relative to the neat polymer (which slows diffusion). Enumerating the role of filler shape and loading on tortuosity is a well-researched topic, and hence this factor, 1+ϕNP2, is a convenient simplification that describes data in the limit of small NP loadings.25 Since the gas selectivity is a ratio of permeabilities, these factors cancel out precisely and predict no changes in gas selectivity on filler addition. While these ideas have been borne out in many hybrids,26,27 some physical polymer/NP mixtures display unexpected enhancements in gas permeability.28-31 At this time there is no clear understanding of when the permeability is increased or decreased, and what precise molecular factors govern this behavior. However, it has been speculated that the spatial dispersion of the NPs in the polymer (e.g., well dispersed vs agglomerated) is critical.

Unfortunately, most inorganic/organic hybrids are immiscible since NPs are typically hydrophilic while polymers are apolar,20-22 leading to uncontrolled and undesired NP agglomeration. A well-explored method to circumvent these problems and ensure good NP dispersion is to chemically graft polymer chains on to the NPs to form one-component polymer-grafted nanoparticles (GNPs).32-34 We focus here on spherical, rigid, inorganic NPs with a well-defined core radius (2nm ≤ Rc ≤ 25nm) uniformly grafted with linear polymer chains. Past work has shown that pure GNP membranes (also termed “matrix-free GNPs”) exhibit controllable increases in CO2 and CH4 permeability with only minor losses in CO2/CH4 selectivity; these findings appear to be in direct contrast to conventional Maxwell theory [Eq. (2)].18,35 In particular, at fixed Rc = 8 nm and areal grafting density, σ ≈ 0.47 chains/nm2, we previously found that the CO2 (and also CH4) permeability varied non-monotonically with graft chain length, with a maximum in the vicinity of Mn ≈ 88 kDa. Based on the unexpected behavior of these materials relative to Maxwell theory, as well as the observed non-monotonic behavior, a comprehensive understanding of the molecular origins of these findings, especially how gas permeability varies with NP core size and polymer graft parameters, is needed.

We begin to understand these results through a newly proposed theory for grafted polymer conformations at high enough σ so that the chains are extended relative to their native Gaussian states.36 Under these conditions the brush has two zones: an inner, dry zone where chain sections are extended relative to their Gaussian conformations, and an outer region where the chain sections are essentially Gaussian and interpenetrated with chain sections from adjacent NPs. Based on this theory, we conjecture that the peak permeability occurs at the chain length where the brush has the highest extension free energy per chain. Under these conditions quasielastic neutron scattering (QENS) experiments indicate that the segmental diffusion of the grafted chains goes through a maximum.37,38 This brush height, hmax=3Rc, is predicted by the theory to be independent of chain chemistry and σ. Thus, the maximum permeability must occur at an apparently universal value of the NP loading, ϕNP, ϕNP,max = [Rc/(Rc + hmax)]3 ≈ 0.049. Inspired by this conclusion, we examine gas permeability enhancements across a variety of Rc, Mn and σ, and find that they do follow an apparently universal dependence as a function of ϕNP for high enough σ. For ϕNP > ϕNP,max (i.e., shorter graft chains at fixed grafting density) we find that the permeability depends inversely on NP volume fraction [1/ϕNP [proportional, variant] Mn]. Since this dependence is very different from the dependence implied by the Maxwell model, Eq. (2), some other factor must play an important role here. The structural model of Midya et al.36 implies that increased chain extension is the source of the increased gas permeability in the GNP layers in this regime relative to the neat polymer. On the other hand, for ϕNP < ϕNP,max (i.e., for longer chains at fixed grafting density) we find that permeability is decreased from the maximum value and follows exp[−C(ϕNP,max − ϕNP)]. These results imply that, in parallel with the structure, the polymer layers of the GNP also behave akin to a two-layer transport medium – the “dry” region in the near vicinity of the NP surface speeds-up gas diffusion. The interpenetration region, which occurs for long enough chains, corresponds to slow gas transport, akin to that in a typical polymer melt. At large chain lengths the gas molecules thus have to cross this slow transport region before they encounter another region of fast transport leading to progressively slower gas diffusion with increasing Mn.

Theoretical Basis:

Midya et al.36 recently proposed a model for the structure of neat GNPs. This theory assumes that a spherical brush in a melt of GNPs can have two regions. The inner dry layer of thickness hdry, with ndry monomers per chain, is comprised of extended chain fragments (Figure 2). This layer is surrounded by an interpenetration region where the remaining segments from each chain, ninter = N − ndry, are mixed with chain sections from adjacent NPs (N is the chain length); these interpenetrated chain segments follow Gaussian statistics. The thickness of the interpenetration zone is hinter. To obtain hdry, hinter, ndry and ninter we use two volume filling conditions:

43π(Rc+hdry)3=43πRc3+Zndryρ
(3a)

43π(Rtot)3=43π(Rc+hdry+hinter2)3=43πRc3+4πRc2σNρ
(3b)

where Z is the number of grafts per NP (4πRc2σ), and ρ is the monomer density (in units of monomers/nm3). We also employ the mass balance condition N = ndry + ninter and the Gaussian chain assumption hinter=ninter12b (b is the statistical segment length) for chain segments in the interpenetration zone. In this framework the NP volume fraction is:

ϕNP=(RcRc+h)3=43πRc343πRc3+ZNρ11+y
(4)

where y=3σNρRc and h = hdry + hinter/2. The solution of Eqs. (3) are a function of a single spherical overcrowding parameter, x,

x=Zπρb2Rtot4σRcρb2ϕNP13
(5)

x ≈ 1 implies that the brush structure is akin to an almost unperturbed melt, while x > 1 implies that the grafted chains are extended. This overcrowding parameter is a more appropriate descriptor of chains grafted to spherical NPs, and is analogous to xplanar=σb2N, used in the case of planar brushes, and frequently also employed in the case of spherical GNPs.33 This formulation, in particular Eq. (3b), yields the extension free energy per chain:

ΔFkBT=32h2Nb2=98xNP1y[(1+y)131]298xNPϕNP131ϕNP[1ϕNP13]2
(6)

An external file that holds a picture, illustration, etc.
Object name is nihms-1829350-f0002.jpg

Schematic representation of the two zones in GNP materials; a dry brush regime where the polymer is locally extended, and an interpenetration zone with chains from neighboring particles being fully mixed with each other. Chain segments in this interpenetration zone follow Gaussian statistics.

This free energy has a maximum value of

ΔFmaxkBT=9σRc2ρb2(23+3)=9xNP8(23+3)0.18xNP
(7)

at

hmax=3Rc
(8)

where:

ymax=3(23+3)Nmax=(23+3)ρRcσ
(9)

independent of σ and ρ. We thus obtain a generic prediction that, at the chain length for maximum extension, the NP volume fraction ϕNP assumes a universal value of:

ϕNP,max=(RcRc+hmax)3=(11+3)30.049
(10)

independent of graft chemistry and other grafting parameters, with the only assumption that the graft density be high enough to ensure the applicability of the two-layer model.

We postulate that the permeability enhancements found in GNP membranes track the chain extensional free energy, with larger gas permeability values obtained for larger ΔF. This framework is particularly attractive since it predicts that the NP volume fraction at which the maximum permeability occurs is independent of any specifics chemistry and of the NPs, with the constraint that the grafting density is high enough to ensure the existence of a dry brush where the two-layer formalism is applicable. This is the central hypothesis of this work, one whose applicability we shall examine critically below.

Results and Discussion

Enhanced gas transport in densely grafted GNPs.

To experimentally validate the theory laid out above, we first consider poly(methyl acrylate) grafted nanoparticles (PMA-GNPs) at fixed σ (≈0.47 chains/nm2) and Rc (=8 nm) with Mn varying from 30 to 200 kDa. Structural analysis using Small Angle X-ray Scattering (SAXS) and Transmission Electron Microscopy (TEM) indicate that the NP cores remain well-dispersed for all Rc, Mn, and σ (see supporting information, Figure S1). CO2 transport in these GNPs, relative to ungrafted poly(methyl acylate) (PMA), is shown in Figure 3A (also shown in Robeson plot format in Figure 1). Based on the Maxwell relation (Eq. 2), all composites should have permeabilities less than that of pure PMA. That is, the data obtained should lie to the left of PMA on the Robeson Plot, Figure 1 or below the dashed line in Figure 3A. This is indeed observed for ungrafted, well-dispersed mixtures of NPs and PMA (filled black circles in Figure 1, also Figure S3 in the supporting information).

An external file that holds a picture, illustration, etc.
Object name is nihms-1829350-f0003.jpg

(A) CO2 permeability in PMA-grafted NPs normalized by the value in the corresponding pure polymer (without NPs) as a function of Mn for a series of grafting densities. Rc=8 nm in parts (A) and (B). The horizontal dashed line represents the neat polymer (i.e., Pϕ/Pn=1). (B) Effects of CO2 diffusion constant and solubility on gas permeability in PMA GNPs at constant areal grafting density (0.47 chains/nm2) as a function of graft Mn. Solubility calculations consider only the volume of the polymer phase. (C) Effect of the NP core size on the permeability enhancement as a function of Mn. The σ ≈ 0.47 chains/nm2 for the larger silica NPs, while it is σ ≈ 0.25 chains/nm2 for the smallest zirconia NP. The grafts are PMA for the two largest NPs while it is PMMA for the smallest. (D) Normalized CO2 permeability of PS-grafted NPs at low graft density, σ ≈ 0.03 chains/nm2, as a function of the volume fraction of the silica NPs (Rc=8 nm). The line is the prediction of the Maxwell model.

However, all grafted NP analogs at this Rc and σ exhibit enhanced CO2 permeabilities relative to that of pure PMA; this result is in agreement with our previous report.18 Indeed, the CO2 permeability of the GNPs increases by roughly an order of magnitude for grafted PMA composites with Mn ≈ 88 kDa. These increases are coupled with slight decreases in CO2/CH4 selectivity.

We next consider PMA-based GNPs of the same Rc, over a range of areal graft densities, σ ≈ 0.11 to 0.66 chains/nm2 (Figure 3A; also shown in Robeson plot format in the supporting information, Figure S4). The normalized PCO2 at each grafting density changes non-monotonically with grafted chain Mn in each case, with a maximum for a grafted chain Mn between 60 and 100 kDa; the location of the maximum appears to be only weakly dependent on σ. However, increasing the σ at a given Mn results in a monotonic increase of the relative CO2 permeability, though this is accompanied with successively larger decreases in CO2/CH4 selectivity. Since our theory predicts that the extensional free energy of the chains is proportional to σ [Eq. (6)], the experiments and theory appear to be consistent if the chain extension free energy indeed underpins permeability enhancements.

We also tested poly(methyl methacrylate) [PMMA] and polyaniline-grafted SiO2 GNPs and find that the enhancements in gas transport properties apparently are independent of polymer chemistry, as predicted by theory (see previous work18,35, see supporting information, Figure S5). Importantly, while each of the GNPs tested shows permeability increases relative to the neat polymer melts, the absolute properties of the GNPs are tied to that of the grafted polymer — e.g., GNPs with grafted PMMA are less permeable but more selective than GNPs based on PMA for comparable chain Mn and σ. These results are reflective of the fact that neat PMMA (without NPs) is less permeable to CO2 (and more selective for CO2/CH4 separations) than neat PMA. When the data are normalized by the pure polymer reference then we get similar trends in all cases, emphasizing that the beneficial gas permeability increases alone are intimately connected primarily to the physics of the grafted polymer “brush”.

These permeability increases can be separated into changes in gas diffusivity (D) and solubility (S) in the framework of the solution diffusion model [Eq. (1)]. Figure 3B clearly shows that the permeability enhancements are almost entirely the result of increased diffusion, while gas solubility is unaffected relative to the neat polymer in all cases (within experimental uncertainty). The negligible solubility changes do not contradict the Maxwell model which suggests that S/Sn = (1 − ϕNP) on the addition of impenetrable NPs [the numerator in Eq. (2)]. Hence, almost all of the interesting variations seen in the permeability data are driven by changes in diffusion, which are much more pronounced in terms of their ϕNP dependence than that anticipated by conventional theories.

Variations in GNP core size.

To investigate the effect of NP size on gas transport, we studied two additional cases: (i) PMA chains densely grafted to Rc = 25 nm silica NPs (σ ≈ 0.47 chains/nm2) and (ii) PMMA chains grafted to Rc = 2 nm ZrO2 NPs (σ ≈ 0.25 chains/nm2, Figure 3C). We first consider Rc = 25 nm silica NPs grafted with PMA. For Mn < 50 kDa the PCO2 values are below that of the pure PMA (Figure 3C), and there is no significant change in CO2/CH4 selectivity (see supporting information, Figure S6). Beyond this molecular weight, the CO2 permeability of the GNP increase monotonically with increasing Mn; however, we do not see a maximum for the molecular weights considered.

Experiments conducted with the Rc = 2 nm ZrO2 NPs grafted with PMMA show similar behavior to the high graft density PMA brushes on the Rc = 8 nm silica NPs discussed above. However, while there is a local maximum in permeability, the graft Mn at which it occurs is significantly smaller than for the larger silica GNPs (Mn ≈ 7-10 kDa vs. Mn ≈ 60-90 kDa for the Rc = 8 nm NPs).

While we only have limited data on variations in NP core size, grafted chain chemistry, Mn and σ, below we will show that the data are supportive of theoretical ideas which appear to suggest that permeability enhancement values only depend on the NP volume fraction. Clearly, however, much more data need to be obtained to place these conclusions on a more solid footing.

Suppressed gas transport in sparsely grafted GNP.

In contrast to the trends discussed above for densely-grafted GNPs, very low graft density materials (PS grafted Rc = 8nm NPs with σ ≈ 0.03 chains/nm2) show very different behavior. In these materials, we observe a suppression of gas permeability regardless of chain Mn, though longer chains (i.e., smaller NP loadings) exhibit less of a suppression (Figure 3D). The trends for both CO2 and CH4 permeability are well-described by the Maxwell equation, suggesting GNPs with a grafting density below a threshold value behave as expected by macroscopic theories where the (impermeable) NPs hinder gas transport. Where the crossover to hindered transport occurs, and why this occurs, are important underpinning questions we shall consider below.

Empirical unifying trends in the high graft density data.

We have conjectured above that the peak permeability occurs at an apparently universal value of the NP volume fraction, namely ϕNP,max = 0.049. Inspired by this prediction, we have re-examined all of our CO2 permeation data (except for the lowest graft density of σ ≈ 0.03 chains/nm2, where the two-layer model does not apply) as functions of ϕNP in Figure 4A; Figure 4B shows the corresponding data for CH4. Note that the GNP permeability data are normalized by the corresponding values in the pure polymer – if the theory is correct then this takes in to account the effects of chain chemistry on gas permeability. Within the relatively large scatter, it is apparent that the data for each gas follow an essentially universal non-monotonic dependence of permeability on ϕNP. The permeability of both gases initially increases with ϕNP for small ϕNP < 0.04, but then it decreases with increasing ϕNP for ϕNP> 0.04. There is a peak in relative permeability in the vicinity of ϕNP ≈ 0.04-0.05, which is close to the predicted value (ϕNP,max = 0.049) where the (unfavorable) chain extension free energy is maximum. While these are satisfying general trends, there are important exceptions to these ideas which we discuss below. Regardless, it is clear that the permeability enhancements in these GNP layers apparently follow a universal dependence on the ϕNP.

An external file that holds a picture, illustration, etc.
Object name is nihms-1829350-f0004.jpg

Permeability enhancements of (A) CO2 and (B) CH4 as a function ϕNP for all GNP membranes discussed in Table 1 except for the lowest grafting density.

We use the two-layer model to calculate the molecular weight at which maximum chain extension occurs and from there the overlap parameter at this chain length, xmax. Table 1 shows that the experimentally studied systems fall into three distinct classes. The last three rows in this Table, with xmax >1, are in the regime where the brushes are extended. These are the systems that are most likely to follow the two-layer model, and indeed they show enhanced CO2 and CH4 permeation behavior when examined as a function of ϕNP (Figure 4).

Table 1:

Quantifying overcrowding. The molecular weight Mn,max corresponds to the chain length at which ϕNP,max ≈ 0.049 for PMA-based GNPs. The parameter xmax corresponds to the value of the overcrowding parameter at this chain length, while xNP corresponds to the overcrowding parameter computed using the NP radius.

Rc (nm)σ (chains/nm2)Mn,max (kDa)xmaxxNP ΔFmaxkBT
20.253300.761.180.21
80.0312250.080.220.04
80.113300.772.080.37
80.47793.258.871.60
80.66564.5612.452.24
250.4724810.1327.44.93

Two other sets of data, the one corresponding to the Rc = 8 nm and σ ≈ 0.11 chains/nm2 and the other to Rc = 2 nm with σ ≈ 0.25 chains/nm2, have xmax ≈0.5-1 but these GNPs have extension free energies less than thermal energy. These systems, which are not at high enough graft density, thus are in a regime where the two-layer model is not an effective descriptor of chain conformations in these grafted layers. While the permeability data for these weakly stretched layers qualitatively follow the trends obtained from the brushes with more overlap, there evidently are differences relative to the higher graft density samples especially at ϕNP > 0.1.

The data at the lowest grafting density σ ≈ 0.03 chains/nm2 demonstrate permeability reductions for all volume fractions up to ϕNP = 0.8 (Figure 3D). The x values for this case are always << 1, regardless of graft chain length. Since the two-layer model is clearly inapplicable for these cases, we expect chains from neighboring NPs to readily interpenetrate. These sparsely grafted NPs behave more akin to star polymers in this context with little chain extension, i.e., the chains assume their unperturbed Gaussian conformations. In these cases, the role of NPs on gas transport follows the Maxwell model. While the two-layer model is thus not applicable to explain the trends observed for this low grafting density, the discussion to this point lends credence to our underpinning picture that chain extension is likely the cause of the permeability increases in this class of GNPs.

We now proceed to understand the non-monotonic behavior seen in Figure 4. We have attempted to use extended Maxwell models as suggested in the literature, separately above and below the maximum permeability: we assume that the dry region has significantly higher permeability than the interpenetrated zone, while the NP core is not permeable. This model does not capture the trends observed in Figure 4, in agreement with previous conclusions of Bocharova and coworkers42.

We thus resorted to a more empirical approach. For ϕNP > ϕNP,max we used a variety of forms, including an exponential, a hyperbolic tangent and a power law to see which one adequately describes the data. The exponential fit (i.e., Ce−ϕNP/−NP,max), for example, only fits data in the mid ϕNP range (i.e., 0.049 < ϕNP < 0.1) but fails at larger ϕNP. Instead, we found that a power law form best described the data for both CO2 and CH4. In particular, the line in Figure 5 for the CO2 is a fit that follows:

PPn=0.31ϕNP
(11)

An external file that holds a picture, illustration, etc.
Object name is nihms-1829350-f0005.jpg

Comparison of experimental permeability increases with scaling predictions for (A) CO2 and (B) CH4. We postulate that chain extension dominates gas transport behavior for ϕNP > 0.05; The interpenetration zone, where chains segments behave akin to a melt, dominates for sufficiently long chains (i.e., ϕNP < 0.049).

While the fit power law dependence of P on ϕNP is slightly different from −1, we do not see any reason to use a more complicated dependence than that illustrated in Eq. (11). The CH4 data in Figure 5 follow the same dependence but with a slightly different prefactor, i.e., 0.41 vs. 0.31 for CO2. Here we note that these numbers have significant uncertainties, of order a factor of 2. However, experimental CH4 permeability enhancements are always significantly higher than the CO2 permeability enhancements and hence these uncertainties must be seen as conservative estimates. This result suggests that the transport of CO2 is increased by as large as a factor of ≈ 10 relative to the neat polymer (without NPs), with this increase occurring in the vicinity of the NP volume fraction where the chain extension free energy is the largest. The increase of NP loading leads to a decrease in permeability, so that for large NP loadings, ϕNP > 0.31, GNP permeabilities are predicted to fall below that of the neat polymer. Verification of this conjecture is found experimentally for large NPs with Rc = 25 nm, σ ≈ 0.47 chains/nm2 and Mn = 7.5 kDa, where for example ϕNP = 0.45 we experimentally find PPn=0.3. Eq. (11) predicts PPn=0.65.

While the form that is represented in Eq. (11) is empirically derived, a few points are noted. First, it is quite different from the Maxwell form, Eq. (2). A deeper analysis of the stretching free energy shows that, for ϕNP [dbl greater-than sign] 0.05, it follows the form which is obtained by Taylor expanding the free energy as a function of ln (ϕNP) about ϕNP=1.

ΔFkBT=alnϕNP+b
(12)

where the constant a is dependent on grafting density, while b=0. While the slope of this curve for 8 nm radii NPs assumes a value of a ≈1 for a grafting density of 0.47 chains/nm2, a ≈−1.5 for a grafting density of 0.66 chains/nm2. If we use this form for the free energy of stretching chains, with the assumption that gas motion is facilitated by the free energy of the extended chains with an Arrhenius dependence, we obtain:

PPnϕNPa
(13)

Since theory predicts that the exponent a is strongly dependent on the grafting density and other chemical details of the grafted chains, more comprehensive data, especially for different NP sizes, are critically necessary to further explore the apparent “universality” that appears evident in Figure 5 at large ϕNP.

Of course, while the fact that Eq (11) predicts PPn=0.31 in the limit of ϕNP → 1 evidently accounts for the reduction of permeability due to the impenetrability of the NPs, this is not captured by a simple model that includes only the chain extension free energy. We therefore conjecture that the blocking and increased tortuosity effects modeled in the traditional approaches, especially how they vary with NP loading, play a secondary role to the variation of chain extension free energy with increasing chain length. Apparently, these effects can be modeled by a constant value of 0.31 for CO2 (0.41 for CH4).

Gas permeabilities for NP loadings smaller than ϕNP,max are slowed relative to the maximum. The CO2 data are fit by the empirical equation: PPn=6.4eϕNP,maxϕNPϕpoly,c where we use the requirement that the PPn must go to unity in the limit where the NP loading goes to zero to define ϕpoly,c=ϕNP,maxln(6.4)0.026 (for CO2). The numerator in the exponential accounts for any chain segments outside the dry zone, i.e., those that are unperturbed. The exponential term thus identifies the polymer fragments that are outside the “dry” zone, i.e., the chain segments in the interpenetration zone, as being responsible for permeability decreases, while the prefactor accounts for the transport enhancement that is afforded by the chain segments in the dry region, Ih we assume to be at its maximum extension. The CH4 data behave similarly, except that the peak permeability enhancements (≈ 10) are higher.

Physical picture:

Referring back to our schematic in Figure 2, we conclude that the gas transport in these media is driven by fact that the polymer layer in the GNP has two regions for long enough chains and high enough grafting densities where x > 1. There is an inner regime of extended chains where gas transport is speeded-up relative to the neat polymer, while the outer region of interpenetrated chains is akin to the neat polymer with slow gas transport. Thus, while the gas motion in the inner regions is facilitated by the presence of extended chain segments, penetrant molecules must pass through the slower melt regions to find other fast regions. The fit data suggest that CO2 transport in the region with the most extended chain segments is ~ 6.4 times faster than in the regions with interpenetrated chain segments. The picture proposed is similar in spirit to the hop-like dynamics executed by gas molecules moving through dense polymer matrices with the additional condition that the heterogeneities in the case of the GNPs are created by differential chain extension in different spatial regions of the polymer brush.

Currently Open Issues:

We emphasize that the analysis above has several salient features that are worth stressing. First, the location of the maximum in the case of CO2 permeability in the Rc = 8 nm NPs with σ ≈ 0.47 chains/nm2 occurs near ϕNP,max = 0.049, as predicted. The data for Rc = 8 nm with σ ≈ 0.66 chains/nm2 show a maximum permeability that evidently is at a slightly lower ϕNP,max = 0.035, while the Rc = 25 nm with σ ≈ 0.47 chains/nm2 data do not show a maximum in our experimental measurements; it is currently not feasible to synthesize GNPs with long enough chains to explore if these Rc = 25 nm systems also display a permeability maximum. In the region of higher ϕNP the data from these three samples overlap. Thus, while there are shared features it is unclear if the maximum permeability always occurs at ϕNP,max = 0.049. We need more data, especially in the vicinity of the maximum to unequivocally resolve this situation.

Similarly, the two other cases, Rc = 2 nm with σ ≈ 0.25 chains/nm2 and Rc = 8 nm with σ ≈ 0.11 chains/nm2, which are too sparsely grafted to be in the strongly stretched brush regime with x > 1, show different behavior. While they appear to follow the increase of PPn with decreasing NP loading, there appears to be a local maximum in the data, Figures 4A and and4B,4B, which occurs at much too high a NP loading, ϕNP, between 0.1 and 0.2, relative to what is predicted by the two-layer model. Extensive future experiments will focus on clearly enunciating the location of this maximum.

CONCLUSIONS

We have investigated the enhanced gas transport properties of matrix-free GNPs. The CO2 (and CH4) transport performance of the GNPs is found to be easily varied through changes in Mn, σ, and NP core size, with the changes relative to the pure polymer being apparently independent of polymer chemistry (i.e., PMA vs. PMMA). The permeability enhancement, relative to the values in the pure polymer, from all brushes above a critical grafting density value, i.e., where chains are extended relative to their Gaussian states, appear to follow an apparently universal behavior when plotted as a function of the NP volume fraction. Of course, the gas permeability in a pure polymer is very sensitive to chain chemistry – what is universal is the relative enhancement in permeability that results when these chains are grafted to a NP. A two-layer brush model allows us to rationalize these results. In particular, our results imply that gases are transported effectively in the inner, extended regime of the brush and that the permeability enhancement is directly related to the extent of chain extension – the higher the extension, the faster the gas transport. In contrast, gas permeability is slow in the outer, melt-like regime where chain sections are interpenetrated with chains from adjacent NPs. We propose that while gas molecules move rapidly in the dry regime, they have to traverse through the interpenetrated regions where transport is significantly slowed. The crossover between these two regimes occurs at an apparently universal value of the NP loading, independent of chain chemistry, grafting density and NP core size and corresponds to the chain length with maximum chain extension. While these results are applicable for high grafting density where the dry regime exists, very sparsely grafted chains lead to a decrease in permeability as expected by macroscopic theories (e.g., the Maxwell model) where the NPs present physical barriers that decrease gas permeability.

Materials and Methods

Spherical silica NPs (50±13 nm and 14±4 nm) were obtained from Nissan and used as received. Spherical zirconia (diameter 4±1 nm) was synthesized using a procedure by Garnweitner et al.39 Poly (methyl acrylate) (PMA), polystyrene (PS) and poly(methyl methacrylate) (PMMA) grafted spherical silica particles were synthesized by surface initiated reversible addition-fragmentation chain transfer polymerization (SI-RAFT) technique.40 The RAFT agent 2-(dodecylthiocarbonothioylthio)propanoic acid (DoPAT) was used for the polymerization. 3-Aminopropyldimethylethoxysilane was purchased from Gelest, Inc. and used as received. DoPAT was purchased from Boron Molecular, Inc. Methyl acrylate (MA, 99%, Acros) and methylmethacrylate (MMA, 99%, Acros) were purified by filtration through an activated basic alumina column. Azobisisobutyronitrile (AIBN) was recrystallized from ethanol twice before use. 1H NMR and 13C NMR (Bruker Avance 300) were conducted using CDCl3 as solvent. Molecular weights and dispersity were determined using gel permeation chromatography (GPC) equipped with a Varian 290-LC pump, a Varian 390-LC refractive index detector, and three Styragel columns (HR1, HR3 and HR4, molecular weight range of 100-5,000, 500-30,000, and 5,000-500,000, respectively). THF was used as eluent for GPC at 30 °C and a flow rate of 1.0 mL/min. The GPC was calibrated with PMMA standards obtained from Polymer Laboratories. The GNP chain Mn’s, grafting densities, and dispersities are shown in the supporting information (Table S1).

The structure of bulk GNP materials was investigated using Small-Angle X-Ray Scattering (SAXS) measurements in tandem with Transmission Electron Microscopy (TEM) imaging. Samples for SAXS were prepared by drying ~50-60 mg of sample mass from a solution of 95/5 THF/Chloroform in atmospheric conditions, and then thermal annealing (see supporting information for annealing procedure) SAXS data was collected from 0.005A−1 to ~2A−1. 2-dimensional SAXS patterns were azimuthally averaged to produce 1-dimensional scattering curves. Data were fit to a single form factor and structure factor using the program SasFit. TEM samples were prepared by drop casing dilute solutions onto Lacey Carbon TEM grids (Electron Micriscopy Services, Hatfield, PA). Drop-casted samples were imaged with a FEI TALOS F200X (ThermoFisher) at the Columbia Nano-Initiative.

Gas Transport:

Steady-state permeabilities of light gases were measured using the constant volume-variable pressure method.40 Films of GNPs were solvent cast into Teflon petri dishes from solutions of known concentration, covered with aluminum foil and allowed to dry in a fume hood at room temperature for 12 hours. The samples were dried at 70 °C under vacuum for at least 24 hours to remove any residual solvent, and then annealed under vacuum above the glass transition of the polymer chains (at 80 °C for PMA materials, 120 °C for PS, and 150 °C for PMMA) for at least 24 hours. The resulting film thicknesses were 80-120 μm. The films were then mounted on to 47 mm brass discs with known inner diameter and supported by Annodisc alumina filter paper (nominal pore size 2 nm) that provides no appreciable resistance to gas transport. Separately we have made ~ 100nm thick films directly on Annodisc alumina filter paper – these two systems provided internally consistent results.

Films are loaded into the closed-volume apparatus and degassed under vacuum (< 20 mTorr) for a minimum of 12 hours. The test apparatus was then isolated from the vacuum and the rate of pressure increase in the downstream chamber of known volume is measured to determine the leak rate of the sample. In all cases the leak rate of ambient gas into the apparatus was less than 10 mTorr/hour. After the leak rate is measured, the system is returned to vacuum and one side of the film is exposed to ≈ 2 atm of the penetrant gas. The permeability was calculated from the steady-state rate of downstream pressure increase: Pi=VdlpuART[(dpddt)ss(dpddt)leak], whereVd is the downstream volume, l is the film thickness, pu/pd is the upstream/downstream pressure, t is the time, A is the gas transport film area, R is the gas constant, and T is the experimental temperature (held at 35 °C). (dpddt)ss is the steady-state rate of pressure increase in the downstream, and (dpddt)leak is the measured leak rate of the apparatus.

Unsteady-state transport experiments on CO2 and CH4 were performed on a Quartz Crystal Microbalance with Dissipation Monitoring (QCM-D) (Biolin Q-Sense, Paramus, NJ). QCM-D experiments were performed on thin (< 5 μm) GNP films. This technqiue provides direct measurements of both the Si and Di. QCM-D sensors were purchased from Biolin Scientific and cleaned using a 5:1:1 mixture of deionized water, ammonium hydroxide, and hydrogen peroxide at 70°C. After cleaning, the sensors were placed in a UV/Ozone chamber for 15 minutes and then loaded into the QCM-D flow cell. The resonant frequency and dissipation of the bare quartz sensor was collected, effectively “taring” it for use as a microbalance. Thin films were spin-cast on the sensors from concentrated GNP solutions (50-80 mg/ml in THF) at low spin speeds (500-800 rpm). Spin-cast sensors were subjected to the same drying and annealing procedure as the bulk film samples described above. After drying and annealing the films were loaded into the QCM-D flow cell and nitrogen was flowed over the crystal to function as a “blank”. No significant mass uptake of the nitrogen gas was detected. The gas flow was then changed from nitrogen to the penetrant (CO2 or CH4) in a stepwise manner, and the change in the crystal resonant frequency and dissipation provides an in-situ measurement of the transient mass uptake of the film.

Additional details on materials and methods, including a detailed description on GNP synthesis, is provided in the supporting information.

Supplementary Material

Supplement

Click here to view.(1000K, docx)

Acknowledgments:

Partial financial support for this research was from the Department of Energy under grant DOE-SC0021272. CRB thanks the National Science Foundation Graduate Research Fellowship Program (Grant # DGE-16-44869).

Footnotes

Competing Interests: The authors declare no competing interests.

Associated Content: All data is available in the main text or the supplementary materials.

References

(1) Park HB; Kamcev J; Robeson LM; Elimelech M; Freeman BD Maximizing The Right Stuff: The Trade-Off Between Membrane Permeability And Selectivity. Science 2017, 356 (6343), 1138–1148. 10.1126/Science.Aab0530. [Abstract] [CrossRef] [Google Scholar]
(2) Murali RS; Sankarshana T; Sridhar S Air Separation By Polymer-Based Membrane. 2013, No. April 2012, 130–186. 10.1080/15422119.2012.686000. [CrossRef] [Google Scholar]
(3) Yampolskii Y Polymeric Gas Separation Membranes. Macromolecules 2012, 45 (8), 3298–3311. 10.1021/Ma300213b. [CrossRef] [Google Scholar]
(4) Baker RW; Lokhandwala K Natural Gas Processing With Membranes : An Overview. 2008, 2109–2121. 10.1021/Ie071083w. [CrossRef] [Google Scholar]
(5) Koros WJ; Zhang C Materials For Next-Generation Molecularly Selective Synthetic Membranes. Nature Materials 2017, 16 (3), 289–297. 10.1038/Nmat4805. [Abstract] [CrossRef] [Google Scholar]
(6) Baker RW; Low BT Gas Separation Membrane Materials: A Perspective. Macromolecules 2014, 47 (20), 6999–7013. 10.1021/MA501488S. [CrossRef] [Google Scholar]
(7) Robeson LM Polymer Membranes For Gas Separation. Current Opinion In Solid State And Materials Science 1999, 4 (6), 549–552. 10.1016/S1359-0286(00)00014-0. [CrossRef] [Google Scholar]
(8) Freeman BD Basis Of Permeability/Selectivity Tradeoff Relations In Polymeric Gas Separation Membranes. Macromolecules 1999, 32 (2), 375–380. 10.1021/Ma9814548. [CrossRef] [Google Scholar]
(9) Robeson LM The Upper Bound Revisited. Journal Of Membrane Science 2008, 320 (1–2), 390–400. 10.1016/J.Memsci.2008.04.030. [CrossRef] [Google Scholar]
(10) Sanders DE; Smith ZP; Guo R; Robeson LM; Mcgrath JE; Paul DR; Freeman BD Energy-Efficient Polymeric Gas Separation Membranes For A Sustainable Future: A Review. POLYMER 2013, 54 (18), 4729–4761. 10.1016/J.Polymer.2013.05.075. [CrossRef] [Google Scholar]
(11) Robeson LM Correlation Of Separation Factor Versus Permeability For Polymeric Membranes. Journal Of Membrane Science 1991, 62 (2), 165–185. 10.1016/0376-7388(91)80060-J. [CrossRef] [Google Scholar]
(12) Park HB; Jung CH; Lee YM; Hill AJ; Pas SJ; Mudie ST; Van Wagner E; Freeman BD; Cookson DJ Polymers With Cavities Tuned For Fast Selective Transport Of Small Molecules And Ions. Science (New York, N.Y.) 2007, 318 (5848), 254–258. 10.1126/Science.1146744. [Abstract] [CrossRef] [Google Scholar]
(13) Jimenez-Solomon MF; Song Q; Jelfs KE; Munoz-Ibanez M; Livingston AG Polymer Nanofilms With Enhanced Microporosity By Interfacial Polymerization. Nature Materials 2016, 15 (7), 760–767. 10.1038/Nmat4638. [Abstract] [CrossRef] [Google Scholar]
(14) Mckeown NB; Budd PM Polymers Of Intrinsic Microporosity (Pims): Organic Materials For Membrane Separations, Heterogeneous Catalysis And Hydrogen Storage. Chemical Society Reviews 2006, 35 (8), 675–683. 10.1039/B600349d. [Abstract] [CrossRef] [Google Scholar]
(15) Maillard D; Kumar SK; Fragneaud B; Kysar JW; Rungta A; Benicewicz BC; Deng H; Brinson LC; Douglas JF Mechanical Properties Of Thin Glassy Polymer Films Filled With Spherical Polymer-Grafted Nanoparticles. Nano Letters 2012, 12 (8), 3909–3914. 10.1021/Nl301792g. [Abstract] [CrossRef] [Google Scholar]
(16) Moll JF; Akcora P; Rungta A; Gong S; Colby RH; Benicewicz BC; Kumar SK Mechanical Reinforcement In Polymer Melts Filled With Polymer Grafted Nanoparticles. Macromolecules 2011, 44 (18), 7473–7477. 10.1021/Ma201200m. [CrossRef] [Google Scholar]
(17) Midya J; Cang Y; Egorov SA; Matyjaszewski K; Bockstaller MR; Nikoubashman A; Fytas G Disentangling The Role Of Chain Conformation On The Mechanics Of Polymer Tethered Particle Materials. Nano Letters 2019, 19 (4), 2715–2722. 10.1021/Acs.Nanolett.9b00817. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
(18) Bilchak CR; Buenning E; Asai M; Zhang K; Durning CJ; Kumar SK; Huang Y; Benicewicz BC; Gidley DW; Cheng S; Sokolov AP; Minelli M; Doghieri F Polymer-Grafted Nanoparticle Membranes With Controllable Free Volume. Macromolecules 2017, 50 (18), 7111–7120. 10.1021/Acs.Macromol.7b01428. [CrossRef] [Google Scholar]
(19) Cheng S; Bocharova V; Belianinov A; Xiong S; Kisliuk A; Somnath S; Holt AP; Ovchinnikova OS; Jesse S; Martin H; Etampawala T; Dadmun M; Sokolov AP Unraveling The Mechanism Of Nanoscale Mechanical Reinforcement In Glassy Polymer Nanocomposites. Nano Letters 2016, 16 (6), 3630–3637. 10.1021/ACS.Nanolett.6B00766. [Abstract] [CrossRef] [Google Scholar]
(20) Rittigstein P; Torkelson JM Polymer-Nanoparticle Interfacial Interactions In Polymer Nanocomposites: Confinement Effects On Glass Transition Temperature And Suppression Of Physical Aging. Journal Of Polymer Science, Part B: Polymer Physics 2006. 10.1002/Polb.20925. [CrossRef] [Google Scholar]
(21) Daniele Cangialosi; Boucher VM; Angel Alegría; Juan Colmenero. Physical Aging In Polymers And Polymer Nanocomposites: Recent Results And Open Questions. Soft Matter 2013, 9 (36), 8619–8630. 10.1039/C3SM51077H. [CrossRef] [Google Scholar]
(22) Smith SJD; Hon Lau C; Mardel JI; Melanie Kitchin; Kristina Konstas; Ladewig BP; Hill MR Physical Aging In Glassy Mixed Matrix Membranes; Tuning Particle Interaction For Mechanically Robust Nanocomposite Films. Journal Of Materials Chemistry A 2016, 4 (27), 10627–10634. 10.1039/C6TA02603F. [CrossRef] [Google Scholar]
(23) Bouma RHB; Checchetti A; Chidichimo G; Drioli E Permeation Through A Heterogeneous Membrane: The Effect Of The Dispersed Phase. Journal Of Membrane Science 1997, 128 (2), 141–149. 10.1016/S0376-7388(96)00303-1. [CrossRef] [Google Scholar]
(24) Maxwell JC A Treatise On Electricity And Magnetism.; Dover Publications, 1954. [Google Scholar]
(25) Barrer RM; Rogers MG; Barrie JA Heterogeneous Membranes - Diffusion In Filled Rubber. Journal Of Polymer Science Part A-General Papers 1963, 1 (8), 2565-. 10.1002/Pol.1963.100010806. [CrossRef] [Google Scholar]
(26) Su NC; Smith ZP; Freeman BD; Urban JJ Size-Dependent Permeability Deviations From Maxwells Model In Hybrid Cross-Linked Poly(Ethylene Glycol)/Silica Nanoparticle Membranes. Chemistry Of Materials 2015, 27 (7), 2421–2429. 10.1021/Cm504463c. [CrossRef] [Google Scholar]
(27) Janes DW; Durning CJ Sorption And Diffusion Of N Alkyl Acetates In Poly ( Methyl Acrylate )/ Silica Nanocomposites. 2013. [Google Scholar]
(28) Merkel TC; Freeman BD; Carolina N; He Z; Pinnau I Sorption , Transport , And Structural Evidence For Enhanced Free Volume In Poly ( 4-Methyl-2-Pentyne )/ Fumed Silica Nanocomposite Membranes. 2003, I (4), 109–123. [Google Scholar]
(29) Merkel TC; Freeman BD; Spontak RJ; He Z; Pinnau I; Meakin P; Hill AJ Ultrapermeable, Reverse-Selective Nanocomposite Membranes. Science (New York, N.Y.) 2002, 296 (5567), 519–522. 10.1126/Science.1069580. [Abstract] [CrossRef] [Google Scholar]
(30) Takahashi S; Paul DR Gas Permeation In Poly ( Ether Imide ) Nanocomposite Membranes Based On Surface-Treated Silica . Part 1 : Without Chemical Coupling To Matrix. 2006, 47, 7519–7534. 10.1016/J.Polymer.2006.08.029. [CrossRef] [Google Scholar]
(31) Ahn J; Chung WJ; Pinnau I; Song J; Du N; Robertson GP; Guiver MD Gas Transport Behavior Of Mixed-Matrix Membranes Composed Of Silica Nanoparticles In A Polymer Of Intrinsic Microporosity (PIM-1). Journal Of Membrane Science 2010, 346 (2), 280–287. 10.1016/J.MEMSCI.2009.09.047. [CrossRef] [Google Scholar]
(32) Fernandes NJ; Koerner H; Giannelis EP; Vaia RA Hairy Nanoparticle Assemblies As One-Component Functional Polymer Nanocomposites: Opportunities And Challenges. MRS Communications; 2013, 3 (1), 13–29. 10.1557/Mrc.2013.9. [CrossRef] [Google Scholar]
(33) Kumar SK; Jouault N; Benicewicz B; Neely T Nanocomposites With Polymer Grafted Nanoparticles. MACROMOLECULES 2013, 46 (9), 3199–3214. 10.1021/Ma4001385. [CrossRef] [Google Scholar]
(34) Srivastava S; Agarwal P; Archer LA Tethered Nanoparticle-Polymer Composites: Phase Stability And Curvature. Langmuir 2012. 10.1021/La2049234. [Abstract] [CrossRef] [Google Scholar]
(35) Bilchak CR; Jhalaria M; Huang Y; Abbas Z; Midya J; Benedetti FM; Parisi D; Egger W; Dickmann M; Minelli M; Doghieri F; Nikoubashman A; Durning CJ; Vlassopoulos D; Jestin J; Smith ZP; Benicewicz BC; Rubinstein M; Leibler L; Kumar SK Tuning Selectivities In Gas Separation Membranes Based On Polymer-Grafted Nanoparticles. ACS Nano 2020. 10.1021/Acsnano.0c07049. [Abstract] [CrossRef] [Google Scholar]
(36) Midya J; Rubinstein M; Kumar SK; Nikoubashman A Structure Of Polymer-Grafted Nanoparticle Melts. ACS Nano 2020, 14 (11), 15505–15516. 10.1021/Acsnano.0c06134. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
(37) Jhalaria M; Huang Y; Ruzicka E; Tyagi M; Zorn R; Zamponi M; Sakai VG; Benicewicz B; Kumar S Activated Transport In Polymer Grafted Nanoparticle Melts. Macromolecules 2021, 54 (14), 6968–6974. 10.1021/ACS.MACROMOL.1C00601. [CrossRef] [Google Scholar]
(38) Jhalaria M; Buenning E; Huang Y; Tyagi M; Zorn R; Zamponi M; García-Sakai V; Jestin J; Benicewicz BC; Kumar SK Accelerated Local Dynamics In Matrix-Free Polymer Grafted Nanoparticles. Physical Review Letters 2019, 123 (15), 158003. 10.1103/Physrevlett.123.158003. [Abstract] [CrossRef] [Google Scholar]
(39) Garnweitner G; Goldenberg LM; Sakhno OV; Antonietti M; Niederberger M; Stumpe J Large-Scale Synthesis Of Organophilic Zirconia Nanoparticles And Their Application In Organic-Inorganic Nanocomposites For Efficient Volume Holography. Small 2007, 3 (9), 1626–1632. 10.1002/Smll.200700075. [Abstract] [CrossRef] [Google Scholar]
(40) Comyn J Introduction To Polymer Permeability And The Mathematics Of Diffusion. In Polymer Permeability; Springer Netherlands: Dordrecht, 1985; Pp 1–10. 10.1007/978-94-009-4858-7_1. [CrossRef] [Google Scholar]
(41) Janes DW; Kim JS; Durning CJ Interval Sorption Of Alkyl Acetates And Benzenes In Poly ( Methyl Acrylate ). 2012. [Google Scholar]
(42) Jeong SP; Kumar R; Genix A-C; Popov I; Li C; Mahurin SM; Hu X; Bras W; Popov I; Sokolov AP; Bocherova V Improving Gas Selectivity in Membranes using Polymer-Grafted Silica Nanoparticles, ACS Appl. Nano. Mater, 2021, 4(6), 5895–5903. 10.1021/acsanm.1c00803. [CrossRef] [Google Scholar]

Citations & impact 


This article has not been cited yet.

Impact metrics

Alternative metrics

Altmetric item for https://www.altmetric.com/details/126556199
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/126556199

Smart citations by scite.ai
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
Explore citation contexts and check if this article has been supported or disputed.
https://scite.ai/reports/10.1021/acs.macromol.1c02367

Supporting
Mentioning
Contrasting
1
14
0

Similar Articles 


Funding 


Funders who supported this work.

Deutsche Forschungsgemeinschaft (1)

Division of Graduate Education (1)

NHLBI NIH HHS (1)

U.S. Department of Energy (1)