You are on page 1of 10

Subscriber access provided by NESMEYANOV INST ORGANOELEMENT

Article
Enzyme Encapsulation in a Porous Hydrogen-Bonded Organic Framework
Weibin Liang, Francesco Carraro, Marcello B Solomon, Stephen G. Bell, Heinz Amenitsch,
Christopher J. Sumby, Nicholas G. White, Paolo Falcaro, and Christian J. Doonan
J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.9b06589 • Publication Date (Web): 19 Aug 2019
Downloaded from pubs.acs.org on August 27, 2019

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a service to the research community to expedite the dissemination
of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in
full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully
peer reviewed, but should not be considered the official version of record. They are citable by the
Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore,
the “Just Accepted” Web site may not include all articles that will be published in the journal. After
a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web
site and published as an ASAP article. Note that technical editing may introduce minor changes
to the manuscript text and/or graphics which could affect content, and all legal disclaimers and
ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or
consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 9 Journal of the American Chemical Society

1
2
3
4
5
6
7 Enzyme Encapsulation in a Porous Hydrogen-Bonded
8 Organic Framework
9
10 Weibin Liang†∇, Francesco Carraro§∇, Marcello B. Solomon§, Stephen G. Bell†, Heinz Amenitsch∥,
11
Christopher J. Sumby*†, Nicholas G. White*‡, Paolo Falcaro*§,†, Christian J. Doonan*†
12
13 † Department of Chemistry and the Centre for Advanced Nanomaterials, The University of Adelaide, Adelaide, South
14 Australia 5005, Australia
15 § Institute of Physical and Theoretical Chemistry, Graz University of Technology, Stremayrgasse 9, Graz 8010, Austria
16
17
∥ Institute of Inorganic Chemistry, Graz University of Technology, Stremayrgasse 9, Graz 8010, Austria
18 ‡ Research School of Chemistry, The Australian National University, Canberra, ACT, Australia
19 ∇ These authors contributed equally to the article
20
21
22 ABSTRACT: Protection of biological assemblies is critical to applications in biotechnology, increasing the durability of enzymes
23 in biocatalysis or potentially stabilizing biotherapeutics during transport and use. Here we show that a porous hydrogen-bonded
24 organic framework (HOF) constructed from water soluble tetra-amidinium (1·Cl4) and tetracarboxylate (2) building blocks can
25 encapsulate and stabilize biomolecules to elevated temperature, proteolytic and denaturing agents, and extend the operable pH
range for catalase activity. The HOF, which readily retains water within its framework structure, can also protect and retain the
26
activity of enzymes such as alcohol oxidase, that are inactive when encapsulated within zeolitic imidazolate framework (ZIF)
27 materials. Such HOF coatings could provide valid alternative materials to ZIFs: they are metal free, possess larger pore apertures,
28 and are stable over a wider, more biologically relevant pH range.
29
30
31
32 INTRODUCTION topologies observed for reported ZIF biocomposites possess
33 Protection of biological assemblies is critical to applications in narrow crystallographic pore apertures (typically ~ 3.4 Å) that
34 biotechnology, such as biocatalysis and therapeutics, which significantly limit the size of substrates that can be used for
often expose enzymes to harsh environments known to have biocatalytic transformations.19, 20 In addition, ZIF-8 can
35
deleterious effects on their native activity.1, 2 One approach to decompose in the presence of certain buffer solutions,21, 22
36 chelating agents,14 and at mildly acidic pH.23, 24 This instability
enhance enzyme stability is to immobilize them within porous
37 can limit the diversity of chemistry that can be explored for
materials. Indeed, research has shown that post-infiltration of
38 enzymes into mesoporous silicas,3 metal-organic frameworks such systems. Thus, to develop this area of porous framework
39 (MOFs),4 and covalent-organic frameworks5 offers a level of chemistry we sought to uncover new materials for the
40 protection from organic solvents and elevated temperatures. encapsulation and protection of biomolecules.
41 However, infiltration of guests into pre-synthesized materials Another class of crystalline materials that may be synthesized
42 has the inherent limitation that the pore apertures of the host via a building block approach are hydrogen-bonded organic
43 material are larger than the diameter of the guest protein frameworks (HOFs).24-26 HOFs are porous network materials
44 molecule. Recently, we and others have reported a different assembled by hydrogen bonding between organic components,
45 approach where biomolecules and assemblies thereof (e.g. and are often highly crystalline and can be prepared in mild
46 enzymes6-11 and viruses12-14) induce the growth of MOFs to conditions. Prototypical systems were first prepared in the
form a porous coating that significantly increases their 1990s,27, 28 while more recent examples have been shown to be
47
durability when exposed to challenging conditions. In this surprisingly stable and/or permanently porous.29-32 While
48
process, the MOF components self-assemble around the HOFs are often incompatible with polar organic solvents such
49 biomolecule encapsulating it within its porous crystalline as DMF and DMSO,25 we have recently reported frameworks
50 matrix. This facile encapsulation method is carried out in prepared from poly-amidinium and poly-carboxylate tectons,
51 aqueous solution at room temperature to ensure minimal which are stable in water and polar solvents, even at 100 °C.33,
52 disruption to the structure of the biomolecule. As a result, the 34 Given these properties, we hypothesized that such HOFs

53 strategy is limited to MOFs that are compatible with these could encapsulate biomolecules to form a new class of porous,
54 synthetic requirements. Zeolitic imidazolate framework-8 crystalline biocomposites. HOFs share analogous synthetic
55 (ZIF-8), and related ZIF materials, have been proven to be design principles with MOFs35, 36 which offers opportunities to
56 excellent candidates for the synthesis of bio-composites.15-18 construct composites of bespoke pore shape, size and
57 Although ZIFs possess many properties that make them functionality. However, they have the advantage of being
favorable for studying biomolecule encapsulation, they present metal free which is desirable for some biological
58
challenges for certain applications. For example, the structure applications.37, 38
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 2 of 9

Here we report the synthesis, characterization and catalytic from elevated temperatures, a proteolytic enzyme (trypsin) and
1 activity of HOF-based biocomposites synthesized from tetra- a chaotropic agent (urea). Also, in the case of FCAT, the HOF
2 amidinium (1·Cl4) and tetracarboxylic acid (H42) building coating extends its maximum activity over a broad pH range
3 blocks (Figure 1 and S1). Throughout this work we refer to (5 to 10). Interestingly, FAOx retains enzymatic activity
4 this biocompatible HOF as BioHOF-1, and enzymes within the HOF; however, the same enzyme shows no activity
5 encapsulated with this HOF as enzyme@BioHOF-1. These when adsorbed on, or encapsulated in ZIF materials. This
6 crystalline composites are stable in water over a wide pH provides a clear example of how HOFs can expand the
range (5-10), polar organic solvents and phosphate buffers. chemistry of enzyme encapsulation in crystalline frameworks.
7
Further, they possess pore dimensions (limiting pore diameter We believe that the control over the topology, pore structure,
8 ca. 6.4 Å) that significantly exceed those found ZIF-based and chemistry offered by HOFs will provide new opportunities
9 materials of sodalite topology (ca. 3.4 Å). We show that the to capitalize on bioactive composites for applications in bio-
10 HOF coating protects encapsulated enzymes (FITC-tagged catalysis and bio-medicine.
11 catalase, FCAT, and FITC-tagged alcohol oxidase, FAOx)
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26 Figure 1. (a) Chemical structures of the BioHOF-1 components. (b) A schematic representation of the synthesis of enzyme@BioHOF-1
27 composites. (c) CLSM images of the as-synthesized FCAT@BioHOF-1 composite and (d) SEM image of FCAT@BioHOF-1 crystals.
28 SEM images of BioHOF-1, FCAT@BioHOF-1, and FAOx@BioHOF-1 are reported in Figure S8.
29 RESULTS AND DISCUSSION adding positively charged 1 to a solution of FCAT followed by
30 addition of negatively charged 2. The resulting solid was
Recently, some of us reported that amidinium and carboxylate
31 functionalized organic molecules could assemble in aqueous separated from the reaction solution, washed three times with
32 solutions to rapidly form porous crystalline networks via deionized water (Figure S19) and stored in water at 4 °C. We
33 charge-assisted hydrogen bonds.33, 34 The biocompatible examined the crystallinity of the product by performing
34 synthesis and stability towards a range of common organic powder X-ray diffraction (PXRD) experiments (Figure S3).
35 solvents (polar and non-polar) led us to hypothesize that these Inspection of the PXRD pattern showed peaks characteristic of
36 novel HOFs could form biocomposites analogous to those BioHOF-1 at 2 values of 8.6°, and 17.4°, indicating that the
37 reported for ZIFs. Accordingly, we selected 1 and 2 (Figure 1 HOF forms in the presence of the enzyme (Figure S3 and S26).
and S1) as building units as they are known to form an open Next, we employed confocal laser scanning microscopy
38
framework structure, here termed BioHOF-1, with limiting (CLSM) to visualize the spatial distribution of FCAT within
39 the HOF-based biocomposites. For comparison, we also
pore dimensions of ca. 6.4 Å (Figure S1 and S2). Initial
40 synthesized FCAT-on-BioHOF-1 by mixing FCAT with pre-
studies targeted the encapsulation of catalase (CAT), a
41 tetrameric enzyme (9.7 × 9.2 × 6.7 nm) that catalyzes the synthesized BioHOF-1 crystals in water. Figures 1c, S4, and
42 splitting of hydrogen peroxide to water and oxygen39 S5 show a clear difference between the CLSM images of
43 (fluorescein-tagged catalase, FCAT, was used herein to FCAT-on-BioHOF-1 and FCAT@BioHOF-1. As expected,
44 facilitate microscopy analysis). Our previous work on ZIF-8 the FCAT molecules were distributed homogeneously within
45 suggested that biocomposite synthesis was facilitated via the the FCAT@BioHOF-1 crystals, which is consistent with
46 accumulation of Zn2+ cations at the protein surface.40 With this enzyme encapsulation. Conversely, the images of FCAT-on-
47 in mind we employed dynamic light scattering (DLS) to BioHOF-1 show greater fluorescence intensity on the crystal
measure the difference in hydrodynamic diameter (Dh) of surfaces. The loading of FCAT in FCAT@BioHOF-1 was
48
FCAT after addition of the amidinium and carboxylate determined to be 6.0 ± 0.5 wt% by ICP-MS.41 The
49 incorporation of FCAT into the composite is also supported by
functionalized building blocks, respectively. Addition of the
50 solid-state UV-visible spectroscopy data that show the Soret
positively charged amidinium link to a solution containing
51 absorption band at 407 nm (π-π*), due to the iron-heme
FCAT (measured zeta potential -15.8  0.1 mV) engendered
52 an dramatic increase in Dh from 10.8 ± 1.0 to 2050 ± 578.4 nm cofactor in CAT (Figure S9).42 Furthermore, the congruence of
53 (Table S1). However, adding the carboxylate-functionalized the Soret absorption peak in free FCAT and FCAT@BioHOF-
54 link did not change the Dh of FCAT. These data support the 1 is indicative that the encapsulation process did not lead to
55 hypothesis that electrostatic attraction between the protein significant structural changes at the enzyme active site.42
56 surface and framework components are important in Gas adsorption isotherm experiments (77K N2) indicated that
57 promoting the initial formation of the HOF biocomposites. To activated samples of FCAT@BioHOF-1 were non-porous to
58 this end, FCAT@BioHOF-1 composites were synthesized by N2. In addition, the parent HOF is also non-porous to N2.
59
60 ACS Paragon Plus Environment
Page 3 of 9 Journal of the American Chemical Society

These data are consistent with our previous results for activity of a control experiment carried out at room
1 amidinum-carboxylate-based HOFs and may be attributed to temperature (RT, 25 ± 1 °C). Free FCAT rapidly lost activity
2 the loss of structurally important water molecules upon from ca. 1.23 µM µgFCAT-1 s-1 at RT to 0.02 µM µgFCAT-1 s-1
3 activation.33 We then turned our efforts to evaluating the after exposure to 60 °C temperatures for 30 min (Figure 2b
4 accessible porosity of the HOF in solution via fluorescein dye and S23). However, the BioHOF-1 coating significantly
5 adsorption. The molecular shape of fluorescein can be enhanced the thermostability of the encapsulated enzyme
6 described as an oblate spheroid with semi-axes of 2 and 7 Å,43 maintaining ca. 79% of its initial activity after 30 min at 60 °C
thus we anticipated that it should diffuse through the pore (Figure 2b).
7
network of the HOF (crystallographically determined pore Next, we sought to establish the activites of encapsulated, and
8 apertures of 6.4 Å, Figure S1). We exposed the HOF crystals
9 surface absorbed, enzyme after exposure to trypsin (a
to fluorescein for 4 hours and then collected CLSM images to proteolytic agent) and urea (a chaotropic agent). Thus, we
10 identify where the dye was spatially located within the crystals. synthesized FCAT@BioHOF-1, FCAT-on-BioHOF-1 and
11 Inspection of the images shows that fluorescein is treated these materials, along with FCAT, with both trypsin
12 homogeneously distributed throughout the HOF (Figure S10), and urea, respectively. Enzymatic activity and structural
13 thus confirming solution accessible porosity. In addition, we characterizations are reported in Figure S24-S29. Trypsin is a
14 carried out analogous solution adsorption experiments for ZIF- proteolytic enzyme that due to its size (approximated as an
15 8, ZIF-90 and MAF-7 which all have reported crystallographic ellipsoid with dimensions of 4.8 × 3.7 × 3.2 nm)46 can only
16 pore network apertures of ~3.4 Å (note that rotation of the access surface bound and not the embedded enzyme.
17 imidazolate rings of ZIF-8 can allow molecules of ~5.0 Å to Subsequent to trypsin exposure, FCAT is completely
18 access the pores).44 Even after soaking in fluorescein solution deactivated whilst the activity of the surface bound FCAT
for 24 hours, the CLSM images show that the dye is located (FCAT-on-BioHOF-1) is significantly reduced (ca. 80% loss
19
only on the surface of each of the ZIF crystals (Figure S10). In of its original biological activity, Figure 2b, S24, and S28).
20 summary, these experiments confirm that the HOF pores are However, FCAT@BioHOF-1 retains 76% of its original
21 accessible to guests in solution and highlight that the larger activity (Figure 2b and S24). These data clearly show that the
22 pore apertures, with respect to ZIF-8, ZIF-90 and MAF-7, enzyme retains activity post-encapsulation and surface
23 allow for the exploration of an expanded number of enzyme adsorption effects are not solely responsible for the protection
24 substrates. observed. Next we exposed FCAT@BioHOF-1, FCAT-on-
25 Next we compared the rate of H2O2 decomposition catalyzed BioHOF-1 and free FCAT to urea; a small molecule that
26 by free FCAT and FCAT@BioHOF-1 in a variety of engenders protein unfolding and loss of activity.47 After
27 conditions. The H2O2 decomposition rate (Vobs) was quantified treating FCAT and FCAT-on-BioHOF-1 with urea a
28 via the FOX assay. As a control experiment, we confirmed significant lost in the activities was observed (ca. 90% and 70%
29 that neat BioHOF-1 crystals do not catalyze the decomposition of the original activity was loss for FCAT and FCAT-on-
of H2O2 (Figure S21). Conversely, when exposed to H2O2, at BioHOF-1, respectively; Figure 2b, S25, and S28). However,
30
room temperature and pH 8 FCAT@BioHOF-1 retained FCAT@BioHOF-1 retains ca. 75% activity (Figure 2b and
31 S25). Given that urea is small enough to diffuse within the
activity (Vobs = 0.27 µM µgFCAT-1 s-1, Figure S22). Given this
32 HOF pore network these data suggests that the micro-
positive result, we assessed the capacity of the HOF coating to
33 protect CAT over the pH range 5-10 as it is known that the environment that houses the enzyme prevents unfolding.48
34 optimal activity for free CAT is at pH of 7-8.45 Figures 2a and Presumably this occurs via tight encapsulation and/or protein-
35 S22 show the optimal activity for the biocomposite is surface interactions.8, 48 Lastly, we performed ten
36 significantly broadened, indeed, >90% of the maximum catalysis/filtration cycles using FCAT@BioHOF-1 to assess
37 activity is retained for FCAT@BioHOF-1 operating over the the catalytic durability of the biocomposite. Figure S30 and
38 biologically-relevant pH range 5-10. To confirm that the HOF S31 show that cycling does not lead to an appreciable
39 retains porosity over a pH range 5-10 via fluorescein dye reduction in enzymatic activity or loss of crystallinity. In
adsorption, HOF crystals were exposed to pH 5 and pH 10 addition, the supernatant of the FCAT@BioHOF-1 composite
40
solutions for 30 min, respectively. Next the respective samples showed negligible enzymatic activity which suggests that
41 enzyme leaching is minimal (Figure S21).
were treated with fluorescein and CLSM images were
42
collected. Figure S11 shows that in both cases the dye
43 adsorbed homogeneously within the pore network indicating
44 that the HOF remains porous after these treatments. Further,
45 1H NMR, PXRD, Raman, UV-visible, and FTIR spectroscopic

46 analyses show that the HOF is structurally stable at low pH


47 values (Figure S12-S17). In contrast, ZIF-8 degrades even in
48 mildly acidic conditions and phosphate buffer solutions
49 (Figure S14-S16).21, 22 These results represent a notable
50 advantage of HOF-based biocomposites for applications where
51 stability is needed in acidic conditions. However, we note that
under extremely basic conditions, pH 14, the HOF materials
52
degrade while ZIF-8 retains structural integrity (Figure S18).
53
54 Many enzymes are also sensitive to elevated temperatures. Figure 2. (a) Relative activity of free FCAT (blue) and
Thus, we assessed the enzymatic activity of free FCAT and FCAT@BioHOF-1 (red) at different pH. Individual data points
55 from (duplicate experiments) are shown as open and closed
FCAT@BioHOF-1 after heating at 60 °C for 30 min. The
56 activities are shown in Figure 2b and S23 relative to the circles.49 (b) Relative activity (%) of free FCAT and
57
58
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 4 of 9

FCAT@BioHOF-1 after thermal treatment (60°C for 30 min), and present. We previously employed this technique to show that
1 exposure to trypsin (2 mg mL-1 for 2 h) and urea (6 M urea for 30 the growth of ZIF-8 was induced by the presence of bovine
2 min). Individual data points from (duplicate experiments) are serum albumin (BSA) in solution and that the BSA@ZIF-8
3 shown as filled and shaded bars.49 structure possessed pores large enough to accommodate the
4 To examine the scope of this encapsulation strategy we protein.8 As such, we used BSA as a model protein in this
5 examined the activites of HOF-based biocomposites of alcohol study so we could compare the data with that obtained for the
6 oxidase (AOx). AOx is an enzyme that consists of eight BSA@ZIF-8. Time-resolved SAXS experiments were
identical sub units (diameter = 20 nm by electron microscopy50 performed using a stop-flow setup54, 55 and the kinetics of the
7
or 15 nm by DLS (Table S1)) and catalyzes the oxidation of nucleation, growth and crystallization of HOF particles (time
8 resolution 100 ms) were monitored in the presence and
9 short aliphatic alcohols to formaldehyde with concomitant
production of H2O2.51 Notably, FAOx@ZIFx and FAOx-on- absence of BSA. Our data show that nuclei formed within the
10 first 100 ms with or without the addition of BSA.
ZIFx (ZIFx = ZIF-8, ZIF-90, and MAF-7) biocomposites
11 synthesized in this work showed no activity (The synthesis, Subsequently, in both cases, the nuclei form amorphous
12 characterization, and catalytic performance of the FAOx/ZIFs particles which crystallize within 100 s. This interpretation is
13 biocomposites are shown in section S8 of the SI (Figure S32- supported by monitoring the intensity of the (110) diffraction
14 S37). We synthesized FAOx@BioHOF-1 (FAOx = peak of BioHOF-1 at 6.1 nm-1 (Figure S45). After 50 s of
15 fluorescein-tagged AOx) composites (Figure 3a and S41) with mixing, the (110) diffraction peak of BioHOF-1 could be
16 an enzyme loading of 2.9 ± 0.6 wt% and then assessed the clearly observed and increased in intensity until it reached a
activity of composite. The CLSM measurments of maximum after ca. 200 s. These experiments suggest that on
17
FAOx@BioHOF-1 (Figure 3a and S6) suggest a homogeneous time scales > 100 ms, HOF growth is not influenced by the
18 presence of BSA and that encapsulation occurs via particle
19 distribution of the FAOx in the HOF crystals; however, in case
of FAOx-on-BioHOF-1 the FAOx localized on the HOF aggregation. SAXS can also be employed to ascertain the
20 presence of a hierarchical pore structure in the HOF-based
surface (Figure S7). FAOx@BioHOF-1 retained ca. 60% of
21 composites. After 15 min of growth we fitted the SAXS
the activity of free FAOx (ca. 3.5 and 2.1 µM µgFAOx-1 min-1
22 for free FAOx and FAOx@BioHOF-1, respectively) (Figure patterns with a hierarchical structural model of the neat HOF
23 S39). The moderately lower observed rate for the encapsulated and BSA@BioHOF-1 (Figure 4 and S46, and S48). The
24 enzyme may result from a variety of factors inherent and BSA@BioHOF-1 data fit showed the presence of mesopores
25 common to immobilization processes.52, 53 Given that with a radius of gyration (Rg) of 4.9 ± 0.5 nm (Figure 4), that
26 FAOx@BioHOF-1 was enzymatically active, we examined are sufficiently large to accommodate isolated BSA molecules
the HOF’s protective capacity when composites were exposed (Rg =3.02 nm56). Conversely, the data for the pure HOF shows
27
to elevated temperature, trypsin and urea. Notably, the an absence of mesopores (Figure S48).
28
29 BioHOF-1 coating offered significant protection to the
30 embedded enzyme compared to free FAOx (Figure 3b). In
summary, the composite maintained ca. 85% of its original
31
activity after heating at 60°C (compared to 20% for the free
32 enzyme); ca. 70% after exposure to trypsin (compared to 14%
33 for the free enzyme); and, ca. 71% after exposure to urea
34 (compared to 15% for the free enzyme) (Figure 3b and S38,
35 S40-S44). These results exemplify how the scope of enzyme
36 encapsulation in crystalline frameworks can be expanded
37 using BioHOF-1.
38
39
40
41 Figure 4. Fitted SAXS patterns together with the single-
42 components of the fit (Power Law and Guinier knee
43 components)57, 58 of BSA@BioHOF-1 after 700 s of growth. Inset
44 shows the gyration radius (arrows) of BSA and the observed
45 mesopore in the BSA@BioHOF-1.
46 CONCLUSIONS
47 Herein, we showed that a water stable amidinium/carboxylate-
48 based HOF can encapsulate biomacromolecules and form a
49 Figure 3. (a) CLSM image of the FAOx@BioHOF-1 (b) Relative porous, crystalline coating that offers protection from
50 activity (%) of free FAOx and FAOx@BioHOF-1 after heating conditions that would typically lead to loss of their native
51 (60, 70, and 80 °C for 10 min) and exposure to trypsin (2 mg mL-1 activity. In contrast to ZIF-8, the HOF-based biocomposites
52 for 2 h) and urea (6 M urea for 30 min). Individual data points possess larger pore apertures and are stable towards acidic pH
53 from (duplicate experiments) are shown as filled and shaded and phosphate buffers. Furthermore, the HOF coating retains
54 bars.49 the activity and protects the enzyme FAOx in challenging
55 conditions, whereas ZIF-based composites of FAOx do not
Lastly, we carried out small angle X-ray scattering (SAXS)
show activity. Accordingly, HOFs represent a new class of
56 experiments to examine the kinetics of protein@BioHOF-1
57 formation and to determine if a hierarchical pore structure was
58
59
60 ACS Paragon Plus Environment
Page 5 of 9 Journal of the American Chemical Society

modular, crystalline materials poised for further exploration in room temperature for 10 min to form solution B. H42 (3 mg)
1 the area of biomolecule protection. was dissolved in 0.95 mL of H2O and 0.05 mL of 1% NH4OH
2 EXPERIMENTAL SECTION to form solution C. Aliquots of solution C (20 µL) were then
3 added to solution B under gentle stirring every 20 s. The
Syntheses
4 mixture was gently stirred for an additional 1 h to ensure the
All chemicals and solvents were purchased from commercial completion of the synthesis. Thereafter, the FAOx@BioHOF-
5
sources and were used as received without further purification. 1 composite was collected by centrifugation and then washed,
6 The tetrahedral amidinium precursor 1·Cl4 was prepared as dispersed, and centrifuged three times in Milli-Q H2O to
7 previously described.33 remove any unreacted precursors and loosely adsorbed FAOx
8 (Figure S19).
Fluorescein-tagged enzymes
9
Fluorescein isothiocyanate (FITC, 0.5 mg) and catalase (CAT; Characterization
10
Sigma-Aldrich, catalase from bovine liver, 2000-5000 units Powder X-ray diffraction (PXRD)
11
mg-1 protein, 40 mg) were dissolved in a carbonate-
12 PXRD patterns were obtained using a D4 ENDEAVOR X-ray
bicarbonate aqueous buffer solution (0.1 M, pH 9.2, 4 mL) and
13 Diffractometer from Bruker. A Co anode was used to produce
left for two hours in darkness at room temperature under
14 Kα radiation (λ = 1.78897 Å). Flat plate diffraction data was
gentle stirring. The FITC-tagged CAT (FCAT) was recovered
collected over the range 2θ = 5–40 °. The PXRD data were
15 by passing the reaction mixture through an Illustra NAP-25
modified by PowDLL Converter (version 2.68.0.0) and
16 column (GE Healthcare Life Sciences, NSW, Australia). The
expressed as the copper-source irradiated patterns (λ =
17 crude FCAT solution was concentrated through a 10 K
1.54056 Å).
18 membrane by centrifugation (4000 rpm for 20 min), followed
by solvent-exchange with ultrapure water. The concentration- Confocal laser scanning microscopy (CLSM)
19
20 solvent-exchange process was repeated two times to ensure the The presence and spatial location of the fluorophore-tagged
buffer salts were completely removed from the solution. biomolecules in (or on) the composites was determined using
21
Thereafter, the concentrated FCAT aqueous solution was CLSM technique (Olympus FV3000 Confocal Laser Scanning
22 passed through a NAP-25 column again to ensure the Microscope, OLYMPUS). The fluorescein-tagged
23 complete removal of unreacted FITC. The obtained FCAT biomolecules were excited at 488 nm and the fluorescence
24 solution was stored in darkness at 4 °C. signal was collected in a window from 495 to 545 nm.
25 A similar method was used to prepare fluorescein-tagged Scanning electron microscopy (SEM)
26 alcohol oxidase (FAOx, Sigma-Aldrich, alcohol oxidase SEM images were collected using a Philips XL30 Field
27 solution from Pichia pastoris, buffered aqueous solution, 10- Emission Scanning Electron Microscope (FESEM). Prior to
28 40 units/mg protein). analysis, the samples were dispersed in ethanol by sonication,
29 Synthesis of BioHOF-1 drop-cast on an aluminum SEM stage (12 mm), and sputter-
30 Amidinium compound 1·Cl4 (4 mg), was dissolved in H2O (1 coated with platinum to form a thin film (5 nm).
31 mL) to form solution A. Carboxylate compound H42 (3 mg) Inductively Coupled Plasma Mass Spectrometer (ICP-MS)
32 was dispersed in Milli-Q H2O (950 µL) , followed by the ICP-MS was performed on an Agilent 8900x QQQ-ICP-MS.
33 addition of aqueous ammonium hydroxide solution (1% v/v, The free enzyme or enzyme@BioHOF-1 composites
34 50 µL) to deprotonate 2 (solution B). Thereafter, solution B (approximately 1 mg) were dispersed in a solution of
35 was added to solution A under stirring conditions at room HNO3/HCl (0.25 mL of 70% HNO3 (Ajax) and 0.25 mL of 37%
36 temperature.34 Precipitates were formed immediately upon HCl (Chem Supply)) and stored in Eppendorf tubes at room
37 mixing (Figure S47). The reaction mixture was left to gently temperature overnight. The mixture was then centrifuged to
38 stir in the dark for 1 h. The BioHOF-1 material was then remove any particulates in the supernatant. Thereafter, the
recovered by centrifugation, and then washed, dispersed, and clear supernatant (0.4 mL) was diluted to a final volume of 5
39
centrifuged three times in Milli-Q H2O to remove any mL with Milli-Q H2O for ICP-MS analysis. The amount of
40 unreacted precursors.
41 enzyme within the sample was calculated according to a
Synthesis of the FCAT@BioHOF-1 biocomposite standard calibration curve for sulfur (prepared from sulfur
42
1·Cl4 (4 mg) was dissolved in H2O (0.5 mL) to form solution standard solution from Sigma-Aldrich, Sulfur Standard for
43 ICP (TraceCERT®, 1000 mg/L S in H2O)).
44 A. An aqueous solution of FCAT (1 mg, 0.5 mL of 2 mg mL-1
stock solution) was added to solution A and stirred at room Synchrotron Small Angle X-ray Scattering
45
temperature for 10 min to form solution B. H42 (3 mg) was Time-resolved SAXS data was collected on the SAXS
46 dissolved in 0.95 mL of H2O and 0.05 mL of 1% NH4OH to
47 beamline at the ELETTRA synchrotron light source.54
form solution C. Solution C was then added dropwise to Operation occurred at a photon energy of 8 keV covering the
48 solution B under stirring. The mixture was then left to gently range of momentum transfer, q = 4π sin(θ)/λ, between 0.1 and
49 stir for another 1 h to ensure the completion of the synthesis. 7.7 nm-1. The kinetics of the MOF nucleation and growth was
50 Thereafter, the FCAT@BioHOF-1 composite was collected by monitored using a commercial stopped flow apparatus SFM-4
51 centrifugation and then washed, dispersed, and centrifuged (Bio-Logic, Grenoble, France) especially designed for
52 three times in Milli-Q H2O to remove any unreacted synchrotron radiation SAXS investigations. Two
53 precursors and loosely adsorbed FCAT (Figure S19). independently controlled syringes were filled with the
54 Synthesis of the FAOx@BioHOF-1 biocomposite carboxylate ligand solution and the amidinium ligand (or
55 1·Cl4 (4 mg) was dissolved in H2O (0.75 mL) to form Solution amidinium ligand + BSA) solution, respectively. Upon
56 A. An aqueous solution of FAOx (0.5 mg, 0.25 mL of 2 mg triggering by the data acquisition system, the two syringe
57 mL-1 stock solution) was added to solution A and stirred at volumes were mixed and injected into a quartz capillary (1
58
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 6 of 9

mm) placed in the X-Ray beam (the mixing/injection process nm under stirring conditions. The reaction rate was determined
1 lasts a few ms). The volume ratio between the two solutions using the maximum linear rate of the increase of A405.
2 and their concentration was set to replicate the conditions used The generation of H2O2 in the system was calculated
3 for the syntheses in batch. according to the formation of oxidized ABTS in the system
4 A total volume of 900 μL was injected for each experiment. (ABTS, extinction coefficient at 405 nm = 36.8 mM-1 cm-1).
5 Images were taken with a time resolution of 100 ms (detector:
6 Pilatus3 1M, Dectris Ltd, Baden, Switzerland; sample to ASSOCIATED CONTENT
7 detector distance: 1260 mm, as determined with a silver Supporting Information. Information related to synthesis of the
8 behenate calibration sample). All experiments were performed biocomposites, additional experimental and analytical detail for
9 at room temperature. The resulting two-dimensional images the biocomposites. This material is available free of charge via the
were radially integrated to obtain a 1D pattern of normalized Internet at http://pubs.acs.org.
10
intensity versus scattering vector q. The background was
11 collected using MilliQ H2O and subtracted as background AUTHOR INFORMATION
12 from the normalized data.
13 Corresponding Authors
Dynamic light scattering (DLS) and zeta potential *Email: christian.doonan@adelaide.edu.au;
14 measurements
15 paolo.falcaro@tugraz.at; nicholas.white@anu.edu.au;
BSA, FCAT, or FAOx (1 mg) was dissolved in 1 mM PBS christopher.sumby@adelaide.edu.au
16 buffer (pH 7.4, 1 ml). Thereafter, 20 µL aqueous solution of 1
17 or 2 (deprotonated using 1% NH4OH) was introduced into the Author Contributions
18 protein solution. DLS and zeta potentials for the pure protein W. L. and F. C. contributed equally to the article. All authors have
19 or protein/ligand samples were obtained on a Malvern given approval to the final version of the manuscript.
20 ZetaSizer dynamic light scattering instrument using a Quartz Funding Sources
21 cuvettes (DTS2145) and folded capillary zeta potential cell
Australian Research Council Discovery Project (DP170103531)
22 (DTS1070), respectively, at 25 °C.
and Discovery Early Career Research Award (DE170100200).
23 Catalytic performance of FCAT and the FCAT@BioHOF- Notes
24 1 composites The authors declare no competing financial interest.
25 The Ferrous Oxidation in Xylenol orange (FOX) assay was
26 applied to quantify the concentration of H2O2 (Figure S20).59 ACKNOWLEDGMENT
27 FCAT or the FCAT@BioHOF-1 composite was added into This work was supported by the Australian Research Council
28 phosphate buffer (100 mM, pH 8, 0.2 mL). Thereafter, H2O2 (DP170103531 and DE170100200). PF acknowledges LP-03.
29 stock solution (1 mM in H2O, 0.15 mL) was added. The
volume of the reaction mixture was adjusted to 1 mL by H2O. REFERENCES
30
The catalyst dosage (based on FCAT) in the enzymatic 1. Bommarius, A. S.; Paye, M. F., Stabilizing biocatalysts.
31
reactions were 2.6 and 2.9 µg for FCAT and Chem. Soc. Rev. 2013, 42, 6534-6565.
32 FCAT@BioHOF-1, respectively (determined by ICP-MS). At 2. Sheldon, R. A.; Brady, D., The limits to biocatalysis:
33 different time intervals, aliquots of the mixtures (50 L) were pushing the envelope. Chem. Commun. 2018, 54, 6088-6104.
34 sampled and mixed with FOX reagent (950 L) in an
3. Hudson, S.; Cooney, J.; Magner, E., Proteins in
35 Mesoporous Silicates. Angew. Chem. Int. Ed. 2008, 47, 8582-8594.
Eppendorf tube and then incubated for at least 30 min at room 4. Drout, R. J.; Robison, L.; Farha, O. K., Catalytic
36 temperature. After incubation, the samples were centrifuged. applications of enzymes encapsulated in metal–organic frameworks.
37 The UV–visible absorbances at 585 nm for the supernatant Coord. Chem. Rev. 2019, 381, 151-160.
38 were recorded to calculate the H2O2 concentration. The 5. Sun, Q.; Fu, C.-W.; Aguila, B.; Perman, J.; Wang, S.;
39 reaction rate (Vobs, µM µgFCAT-1 s-1) is defined as the initial Huang, H.-Y.; Xiao, F.-S.; Ma, S., Pore Environment Control and
H2O2 decomposition velocity of the enzymatic assay. Enhanced Performance of Enzymes Infiltrated in Covalent Organic
40 Frameworks. J. Am. Chem. Soc. 2018, 140, 984-992.
41 Catalytic performance of FAOx and FAOx@BioHOF-1 6. Lyu, F.; Zhang, Y.; Zare, R. N.; Ge, J.; Liu, Z., One-Pot
42 composites Synthesis of Protein-Embedded Metal–Organic Frameworks with
43 The enzymatic activity of FAOx and the FAOx@BioHOF-1 Enhanced Biological Activities. Nano Lett. 2014, 14, 5761-5765.
44 composite were measured according to a modified protocol 7. Shieh, F.-K.; Wang, S.-C.; Yen, C.-I.; Wu, C.-C.; Dutta, S.;
Chou, L.-Y.; Morabito, J. V.; Hu, P.; Hsu, M.-H.; Wu, K. C. W.;
45 from Sigma-Aldrich. One tablet of 2,2′-azino-bis-(3- Tsung, C.-K., Imparting Functionality to Biocatalysts via Embedding
46 ethylbenzothiazoline-6-sulfonic acid) (ABTS, Sigma-Aldrich, Enzymes into Nanoporous Materials by a de Novo Approach: Size-
47 10 mg substrate per tablet) was dissolved in potassium Selective Sheltering of Catalase in Metal–Organic Framework
48 phosphate buffer (pH 7.5, 0.1 M, 10 mL) to form solution A. Microcrystals. J. Am. Chem. Soc. 2015, 137, 4276-4279.
Oxygen gas was bubbled through solution A for ~5 minutes 8. Liang, K.; Ricco, R.; Doherty, C. M.; Styles, M. J.; Bell, S.;
49 Kirby, N.; Mudie, S.; Haylock, D.; Hill, A. J.; Doonan, C. J.; Falcaro,
before use. Solution B (~250 units mL-1 of peroxidase solution)
50 was prepared by dissolving 1.7 mg of HRP (Sigma-Aldrich, P., Biomimetic mineralization of metal-organic frameworks as
51 148 units mg-1) in Milli-Q H2O (1 mL). In a typical assay test, protective coatings for biomacromolecules. Nat. Commun. 2015, 6,
52 7240.
solution A (1 mL), Milli-Q H2O (1.8 mL), solution B (0.01
9. Liang, K.; Coghlan, C. J.; Bell, S. G.; Doonan, C. J.;
53 mL), and aqueous methanol solution (methanol (0.1 mL) in Falcaro, P., Enzyme encapsulation in zeolitic imidazolate frameworks:
54 Milli-Q H2O (1 mL), 0.1 mL) were mixed in a cuvette (light a comparison between controlled co-precipitation and biomimetic
55 path = 1 cm) under magnetic stirring. Thereafter, free FAOx or mineralisation. Chem. Commun. 2016, 52, 473-476.
56 FAOx@BioHOF-1 stock solution (0.1 mL) was introduced 10. Chen, W.-H.; Vázquez-González, M.; Zoabi, A.; Abu-
57 and the reaction mixture was monitored continuously at 405 Reziq, R.; Willner, I., Biocatalytic cascades driven by enzymes
58
59
60 ACS Paragon Plus Environment
Page 7 of 9 Journal of the American Chemical Society

encapsulated in metal–organic framework nanoparticles. Nat. Catal. networks with large chambers. J. Am. Chem. Soc. 1991, 113, 4696-
1 2018, 1, 689-695. 4698.
2 11. Liu, J.; Liu, T.; Du, P.; Zhang, L.; Lei, J., Metal–Organic 28. Russell, V. A.; Evans, C. C.; Li, W.; Ward, M. D.,
Framework (MOF) Hybrid as a Tandem Catalyst for Enhanced Nanoporous Molecular Sandwiches: Pillared Two-Dimensional
3 Therapy against Hypoxic Tumor Cells. Angew. Chem. Int. Ed. 2019, Hydrogen-Bonded Networks with Adjustable Porosity. Science 1997,
4 58, 7808-7812. 276, 575-579.
5 12. Li, S.; Dharmarwardana, M.; Welch, R. P.; Ren, Y.; 29. Yang, W.; Greenaway, A.; Lin, X.; Matsuda, R.; Blake, A.
6 Thompson, C. M.; Smaldone, R. A.; Gassensmith, J. J., Template- J.; Wilson, C.; Lewis, W.; Hubberstey, P.; Kitagawa, S.; Champness,
Directed Synthesis of Porous and Protective Core–Shell N. R.; Schröder, M., Exceptional Thermal Stability in a
7
Bionanoparticles. Angew. Chem. Int. Ed. 2016, 55, 10691-10696. Supramolecular Organic Framework: Porosity and Gas Storage. J.
8 13. Li, S.; Dharmarwardana, M.; Welch, R. P.; Benjamin, C. Am. Chem. Soc. 2010, 132, 14457-14469.
9 E.; Shamir, A. M.; Nielsen, S. O.; Gassensmith, J. J., Investigation of 30. Li, P.; He, Y.; Guang, J.; Weng, L.; Zhao, J. C.-G.; Xiang,
10 Controlled Growth of Metal–Organic Frameworks on Anisotropic S.; Chen, B., A Homochiral Microporous Hydrogen-Bonded Organic
11 Virus Particles. ACS Appl. Mater. Interfaces 2018, 10, 18161-18169. Framework for Highly Enantioselective Separation of Secondary
14. Luzuriaga, M. A.; Welch, R. P.; Dharmarwardana, M.; Alcohols. J. Am. Chem. Soc. 2014, 136, 547-549.
12 Benjamin, C. E.; Li, S.; Shahrivarkevishahi, A.; Popal, S.; Tuong, L. 31. Li, P.; He, Y.; Zhao, Y.; Weng, L.; Wang, H.; Krishna, R.;
13 H.; Creswell, C. T.; Gassensmith, J. J., Enhanced Stability and Wu, H.; Zhou, W.; O'Keeffe, M.; Han, Y.; Chen, B., A Rod-Packing
14 Controlled Delivery of MOF-Encapsulated Vaccines and Their Microporous Hydrogen-Bonded Organic Framework for Highly
15 Immunogenic Response In Vivo. ACS Appl. Mater. Interfaces 2019, Selective Separation of C2H2/CO2 at Room Temperature. Angew.
11, 9740-9746. Chem. Int. Ed. 2015, 54, 574-577.
16 15. Wu, X.; Hou, M.; Ge, J., Metal–organic frameworks and 32. Yamagishi, H.; Sato, H.; Hori, A.; Sato, Y.; Matsuda, R.;
17 inorganic nanoflowers: a type of emerging inorganic crystal Kato, K.; Aida, T., Self-assembly of lattices with high structural
18 nanocarrier for enzyme immobilization. Catal. Sci. Technol. 2015, 5, complexity from a geometrically simple molecule. Science 2018, 361,
19 5077-5085. 1242-1246.
20 16. Doonan, C.; Riccò, R.; Liang, K.; Bradshaw, D.; Falcaro, 33. Morshedi, M.; Thomas, M.; Tarzia, A.; Doonan, C. J.;
P., Metal–Organic Frameworks at the Biointerface: Synthetic White, N. G., Supramolecular anion recognition in water: synthesis
21 Strategies and Applications. Acc. Chem. Res. 2017, 50, 1423-1432. of hydrogen-bonded supramolecular frameworks. Chem. Sci. 2017, 8,
22 17. Riccò, R.; Liang, W.; Li, S.; Gassensmith, J. J.; Caruso, F.; 3019-3025.
23 Doonan, C.; Falcaro, P., Metal–Organic Frameworks for Cell and 34. Boer, S., Morshedi, M., Tarzia, A, Doonan, C. J. , White,
24 Virus Biology: A Perspective. ACS Nano 2018, 12, 13-23. N. G., Molecular Tectonics: node-and-linker building block approach
18. Du, Y.; Gao, J.; Zhou, L.; Ma, L.; He, Y.; Zheng, X.; to a family of hydrogen bonded frameworks. Chem. Eur. J. 2019, 25,
25 Huang, Z.; Jiang, Y., MOF-Based Nanotubes to Hollow Nanospheres 10006-10012.
26 through Protein-Induced Soft-Templating Pathways. Adv. Sci. 2019, 35. Zhou, H.-C.; Long, J. R.; Yaghi, O. M., Introduction to
27 6, 1801684. Metal–Organic Frameworks. Chem. Rev. 2012, 112, 673-674.
28 19. Eum, K.; Jayachandrababu, K. C.; Rashidi, F.; Zhang, K.; 36. Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O.
Leisen, J.; Graham, S.; Lively, R. P.; Chance, R. R.; Sholl, D. S.; M., The Chemistry and Applications of Metal-Organic Frameworks.
29
Jones, C. W.; Nair, S., Highly Tunable Molecular Sieving and Science 2013, 341, 1230444.
30 Adsorption Properties of Mixed-Linker Zeolitic Imidazolate 37. Tamames-Tabar, C.; Cunha, D.; Imbuluzqueta, E.; Ragon,
31 Frameworks. J. Am. Chem. Soc. 2015, 137, 4191-4197. F.; Serre, C.; Blanco-Prieto, M. J.; Horcajada, P., Cytotoxicity of
32 20. Zhang, H.; Hou, J.; Hu, Y.; Wang, P.; Ou, R.; Jiang, L.; nanoscaled metal–organic frameworks. J. Mater. Chem. B 2014, 2,
33 Liu, J. Z.; Freeman, B. D.; Hill, A. J.; Wang, H., Ultrafast selective 262-271.
transport of alkali metal ions in metal organic frameworks with 38. Grall, R.; Hidalgo, T.; Delic, J.; Garcia-Marquez, A.;
34 subnanometer pores. Sci. Adv. 2018, 4, eaaq0066. Chevillard, S.; Horcajada, P., In vitro biocompatibility of mesoporous
35 21. Luzuriaga, M. A.; Benjamin, C. E.; Gaertner, M. W.; Lee, metal (III; Fe, Al, Cr) trimesate MOF nanocarriers. J. Mater. Chem.
36 H.; Herbert, F. C.; Mallick, S.; Gassensmith, J. J., ZIF-8 degrades in B 2015, 3, 8279-8292.
37 cell media, serum, and some—but not all—common laboratory 39. Fita, I.; Rossmann, M. G., The NADPH binding site on beef
buffers. Supramol. Chem. 2019, DOI: liver catalase. Proc. Natl. Acad. Sci. 1985, 82, 1604-1608.
38 40. Maddigan, N. K.; Tarzia, A.; Huang, D. M.; Sumby, C. J.;
10.1080/10610278.2019.1616089.
39 22. Velásquez-Hernández, M. d. J.; Ricco, R.; Carraro, F.; Bell, S. G.; Falcaro, P.; Doonan, C. J., Protein surface
40 Limpoco, F. T.; Linares-Moreau, M.; Leitner, E.; Wiltsche, H.; functionalisation as a general strategy for facilitating biomimetic
41 Rattenberger, J.; Schröttner, H.; Frühwirt, P.; Stadler, E. M.; mineralisation of ZIF-8. Chem. Sci. 2018, 9, 4217-4223.
42 Gescheidt, G.; Amenitsch, H.; Doonan, C. J.; Falcaro, P., 41. Pröfrock, D.; Prange, A., Inductively Coupled Plasma–
Degradation of ZIF-8 in phosphate buffered saline media. Mass Spectrometry (ICP-MS) for Quantitative Analysis in
43 CrystEngComm 2019, 21, 4538-4544. Environmental and Life Sciences: A Review of Challenges, Solutions,
44 23. Sun, C.-Y.; Qin, C.; Wang, X.-L.; Yang, G.-S.; Shao, K.- and Trends. Appl. Spectrosc. 2012, 66, 843-868.
45 Z.; Lan, Y.-Q.; Su, Z.-M.; Huang, P.; Wang, C.-G.; Wang, E.-B., 42. Dong, X.; Fan, Y.; Yang, P.; Kong, J.; Li, D.; Miao, J.;
46 Zeolitic imidazolate framework-8 as efficient pH-sensitive drug Hua, S.; Hu, C., Ultraviolet-visible (UV-Vis) and fluorescence
delivery vehicle. Dalton Trans. 2012, 41, 6906-6909. spectroscopic investigation of the interactions of ionic liquids and
47 catalase. Appl. Spectrosc. 2016, 70, 1851-1860.
24. Luo, J.; Wang, J.-W.; Zhang, J.-H.; Lai, S.; Zhong, D.-C.,
48 Hydrogen-bonded organic frameworks: design, structures and 43. Pu, Y.; Wang, W.; Dorshow, R. B.; Alfano, R. R.,
49 potential applications. CrystEngComm 2018, 20, 5884-5898. Picosecond polarization spectroscopy of fluorescein attached to
50 25. Lin, R.-B.; He, Y.; Li, P.; Wang, H.; Zhou, W.; Chen, B., different molecular volume polymer influenced by rotational motion.
Multifunctional porous hydrogen-bonded organic framework SPIE: 2012; Vol. 8258.
51
materials. Chem. Soc. Rev. 2019, 48, 1362-1389. 44. Verploegh, R. J.; Nair, S.; Sholl, D. S., Temperature and
52 26. Hisaki, I.; Xin, C.; Takahashi, K.; Nakamura, T., Designing Loading-Dependent Diffusion of Light Hydrocarbons in ZIF-8 as
53 Hydrogen-Bonded Organic Frameworks (HOFs) with Permanent Predicted Through Fully Flexible Molecular Simulations. J. Am.
54 Porosity. Angew. Chem. Int. Ed. 2019, 58, 11160-11170. Chem. Soc. 2015, 137, 15760-15771.
55 27. Simard, M.; Su, D.; Wuest, J. D., Use of hydrogen bonds to 45. Chance, B., Effect of pH upon the reaction kinetics of the
control molecular aggregation. Self-assembly of three-dimensional enzyme-substrate compounds of catalse. J. Biol. Chem. 1952, 194,
56 471-481.
57
58
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 8 of 9

46. Saha, B.; Saikia, J.; Das, G., Correlating enzyme density, 54. Amenitsch, H.; Rappolt, M.; Kriechbaum, M.; Mio, H.;
1 conformation and activity on nanoparticle surfaces in highly Laggner, P.; Bernstorff, S., First performance assessment of the
2 functional bio-nanocomposites. Analyst 2015, 140, 532-542. small-angle X-ray scattering beamline at ELETTRA. J. Synchrotron
47. Bennion, B. J.; Daggett, V., The molecular basis for the Rad. 1998, 5, 506-508.
3 chemical denaturation of proteins by urea. Proc. Natl. Acad. Sci. 55. Grillo, I., Applications of stopped-flow in SAXS and SANS.
4 2003, 100, 5142-5147. Curr. Opin. Colloid Interface Sci. 2009, 14, 402-408.
5 48. Liao, F.-S.; Lo, W.-S.; Hsu, Y.-S.; Wu, C.-C.; Wang, S.-C.; 56. Pernot, P.; Round, A.; Barrett, R.; De Maria Antolinos, A.;
6 Shieh, F.-K.; Morabito, J. V.; Chou, L.-Y.; Wu, K. C. W.; Tsung, C.- Gobbo, A.; Gordon, E.; Huet, J.; Kieffer, J.; Lentini, M.; Mattenet,
K., Shielding against Unfolding by Embedding Enzymes in Metal–
7 M.; Morawe, C.; Mueller-Dieckmann, C.; Ohlsson, S.; Schmid, W.;
Organic Frameworks via a de Novo Approach. J. Am. Chem. Soc. Surr, J.; Theveneau, P.; Zerrad, L.; McSweeney, S., Upgraded ESRF
8 2017, 139, 6530-6533.
BM29 beamline for SAXS on macromolecules in solution. J.
9 49. Cumming, G.; Fidler, F.; Vaux, D. L., Error bars in
Synchrotron Rad. 2013, 20, 660-664.
10 experimental biology. The Journal of Cell Biology 2007, 177, 7-11.
50. Zhang, H.; Loovers, H. M.; Xu, L.-Q.; Wang, M.; Rowling, 57. Aragón, S. R.; Pecora, R., Theory of dynamic light
11 scattering from polydisperse systems. J. Chem. Phys. 1976, 64, 2395-
P. J. E.; Itzhaki, L. S.; Gong, W.; Zhou, J.-M.; Jones, G. W.; Perrett,
12 S., Alcohol oxidase (AOX1) from Pichia pastoris is a novel inhibitor 2404.
13 of prion propagation and a potential ATPase. Mol. Microbiol. 2009, 58. Kotlarchyk, M.; Chen, S. H., Analysis of small angle
14 71, 702-716. neutron scattering spectra from polydisperse interacting colloids. J.
51. Ozimek, P.; Veenhuis, M.; van der Klei, I. J., Alcohol Chem. Phys. 1983, 79, 2461-2469.
15 59. Ou, P.; Wolff, S. P., A discontinuous method for catalase
oxidase: A complex peroxisomal, oligomeric flavoprotein. FEMS
16 Yeast Res. 2005, 5, 975-983. determination at ‘near physiological’ concentrations of H2O2 and its
17 52. Hanefeld, U.; Gardossi, L.; Magner, E., Understanding application to the study of H2O2 fluxes within cells. J. Biochem.
18 enzyme immobilisation. Chem. Soc. Rev. 2009, 38, 453-468. Biophys. Methods 1996, 31, 59-67.
19 53. Küchler, A.; Yoshimoto, M.; Luginbühl, S.; Mavelli, F.;
20 Walde, P., Enzymatic reactions in confined environments. Nat.
Nanotechnol. 2016, 11, 409.
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 9 of 9 Journal of the American Chemical Society

TOC
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 9
60 ACS Paragon Plus Environment

You might also like