nucleation-elongation dynamics of two-dimensional …

46
doi.org/10.26434/chemrxiv.8309339.v1 Nucleation-Elongation Dynamics of Two-Dimensional Covalent Organic Frameworks Haoyuan Li, Austin Evans, Ioannina Castano, Michael Strauss, William Dichtel, Jean-Luc Bredas Submitted date: 21/06/2019 Posted date: 24/06/2019 Licence: CC BY-NC-ND 4.0 Citation information: Li, Haoyuan; Evans, Austin; Castano, Ioannina; Strauss, Michael; Dichtel, William; Bredas, Jean-Luc (2019): Nucleation-Elongation Dynamics of Two-Dimensional Covalent Organic Frameworks. ChemRxiv. Preprint. Homogeneous two-dimensional (2D) polymerization is a poorly understood process in which topologically planar monomers react to form planar macromolecules, often termed 2D covalent organic frameworks (COFs). While these COFs have traditionally been limited to weakly crystalline aggregated powders, they were recently grown as micron-sized single crystals by temporally resolving the growth and nucleation processes. Here, we present a quantitative analysis of the nucleation and growth rates of 2D COFs via kinetic Monte Carlo (KMC) simulations, which show that nucleation and growth have second-order and first-order dependences on monomer concentration, respectively. The computational results were confirmed experimentally by systematic measurements of COF nucleation and growth rates performed via in situ X-ray scattering, which validated the respective monomer concentration dependences of the nucleation and elongation processes. A major consequence is that there exists a threshold monomer concentration below which growth dominates over nucleation. Our computational and experimental findings rationalize recent empirical observations that, in the formation of 2D COF single crystals, growth dominates over nucleation when monomers are added slowly, so as to limit their steady-state concentration. This mechanistic understanding of the nucleation and growth processes will inform the rational control of polymerization in two dimensions and ultimately enable access to high-quality samples of designed two-dimensional polymers. File list (2) download file view on ChemRxiv COFNucleationGrowth_062119_ChemRxiv.pdf (0.90 MiB) download file view on ChemRxiv COFNucleationGrowth-SI-062119_ChemRxiv.pdf (3.59 MiB)

Upload: others

Post on 02-Dec-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

doi.org/10.26434/chemrxiv.8309339.v1

Nucleation-Elongation Dynamics of Two-Dimensional Covalent OrganicFrameworksHaoyuan Li, Austin Evans, Ioannina Castano, Michael Strauss, William Dichtel, Jean-Luc Bredas

Submitted date: 21/06/2019 • Posted date: 24/06/2019Licence: CC BY-NC-ND 4.0Citation information: Li, Haoyuan; Evans, Austin; Castano, Ioannina; Strauss, Michael; Dichtel, William;Bredas, Jean-Luc (2019): Nucleation-Elongation Dynamics of Two-Dimensional Covalent OrganicFrameworks. ChemRxiv. Preprint.

Homogeneous two-dimensional (2D) polymerization is a poorly understood process in which topologicallyplanar monomers react to form planar macromolecules, often termed 2D covalent organic frameworks(COFs). While these COFs have traditionally been limited to weakly crystalline aggregated powders, theywere recently grown as micron-sized single crystals by temporally resolving the growth and nucleationprocesses. Here, we present a quantitative analysis of the nucleation and growth rates of 2D COFs via kineticMonte Carlo (KMC) simulations, which show that nucleation and growth have second-order and first-orderdependences on monomer concentration, respectively. The computational results were confirmedexperimentally by systematic measurements of COF nucleation and growth rates performed via in situ X-rayscattering, which validated the respective monomer concentration dependences of the nucleation andelongation processes. A major consequence is that there exists a threshold monomer concentration belowwhich growth dominates over nucleation. Our computational and experimental findings rationalize recentempirical observations that, in the formation of 2D COF single crystals, growth dominates over nucleationwhen monomers are added slowly, so as to limit their steady-state concentration. This mechanisticunderstanding of the nucleation and growth processes will inform the rational control of polymerization in twodimensions and ultimately enable access to high-quality samples of designed two-dimensional polymers.

File list (2)

download fileview on ChemRxivCOFNucleationGrowth_062119_ChemRxiv.pdf (0.90 MiB)

download fileview on ChemRxivCOFNucleationGrowth-SI-062119_ChemRxiv.pdf (3.59 MiB)

1

Nucleation-Elongation Dynamics of Two-

Dimensional Covalent Organic Frameworks

Haoyuan Lia§, Austin M. Evansb§, Ioannina Castanob, Michael J. Straussb, William R. Dichtelb*

and Jean-Luc Bredasa*

aSchool of Chemistry and Biochemistry, Center for Organic Photonics and Electronics (COPE),

Georgia Institute of Technology, Atlanta, Georgia 30332-0400

bDepartment of Chemistry, Northwestern University, Evanston, IL 60208, USA.

§These authors contributed equally to this work.

Corresponding Authors

*[email protected]

*[email protected]

2

ABSTRACT

Homogeneous two-dimensional (2D) polymerization is a poorly understood process in which

topologically planar monomers react to form planar macromolecules, often termed 2D covalent

organic frameworks (COFs). While these COFs have traditionally been limited to weakly

crystalline aggregated powders, they were recently grown as micron-sized single crystals by

temporally resolving the growth and nucleation processes. Here, we present a quantitative analysis

of the nucleation and growth rates of 2D COFs via kinetic Monte Carlo (KMC) simulations, which

show that nucleation and growth have second-order and first-order dependences on monomer

concentration, respectively. The computational results were confirmed experimentally by

systematic measurements of COF nucleation and growth rates performed via in situ X-ray

scattering, which validated the respective monomer concentration dependences of the nucleation

and elongation processes. A major consequence is that there exists a threshold monomer

concentration below which growth dominates over nucleation. Our computational and

experimental findings rationalize recent empirical observations that, in the formation of 2D COF

single crystals, growth dominates over nucleation when monomers are added slowly, so as to limit

their steady-state concentration. This mechanistic understanding of the nucleation and growth

processes will inform the rational control of polymerization in two dimensions and ultimately

enable access to high-quality samples of designed two-dimensional polymers.

3

TOC GRAPHICS

KEYWORDS: Covalent organic frameworks (COFs); two-dimensional (2D) materials; kinetic

Monte Carlo simulations; two-dimensional polymer networks; in situ X-ray diffraction

4

Introduction

Two-dimensional (2D) polymerization is an emerging frontier in polymer science, which promises

the development of rationally designed 2D materials.1 2D covalent organic frameworks (COFs)

are a class of 2D polymers that are permanently porous, atomically regular, synthetically modular,

and structurally versatile.2-14 These properties have inspired explorations of COFs for applications

such as catalysis, molecular separations, energy storage, and optoelectronic devices.15-34 However,

their small crystalline domain sizes (< 100 nm) and empirical polymerization processes have

impeded the study of their intrinsic properties. It has been suggested that the insufficient materials

quality of the 2D COFs obtained to-date is likely due to uncontrolled nucleation-elongation

processes in the reaction mixture.35-38 However, characterizing the small, polydisperse, constantly

evolving particles present in the early stages of COF polymerization remains a demanding

experimental task. From a theoretical perspective, using an analytical approach to describe the

nucleation and growth rates is also daunting because of the presence of many simultaneous growth

processes and pathways.39 Furthermore, the application of computational methods such as

molecular dynamics (MD) simulations provides only limited information because of the short time

scales that are accessible.40

Recently, Smith et al. demonstrated a novel synthetic method that prevents agglomeration of COF

particles and instead results in a stable colloidal suspension.41 This approach enabled the temporal

resolution of the growth and nucleation processes and the synthesis of several single-crystalline

2D COF colloids.42 However, the underlying mechanism remains poorly understood, which stands

as a barrier to the general application of this approach. Here, we use kinetic Monte Carlo (KMC)

simulations to study the 2D COF nucleation and elongation rates as a function of initial monomer

5

concentration. Based on the simulation results, we are able to reach a quantitative understanding

of the nucleation and growth rates, which we further validate via in situ X-ray scattering

measurements. Overall our results allow us to rationalize why reducing the steady-state monomer

concentration suppresses nucleation and favors crystallite growth, the strategy that has been

successfully used in the fabrication of micron-sized 2D COF crystals.42 This kinetic model will

inspire other strategies to temporally resolve nucleation and growth. Similar control of initiation

and propagation in linear polymerization has revolutionized precision polymer synthesis, and its

further development in two-dimensional systems is key to fully realizing the potential of this

longstanding missing topology in macromolecular science.

Results and Discussion

To probe the general dynamics of 2D COF polymerization, we focus on the boronate ester-linked

COF-5, which is a frequently used model system in the study of 2D COFs.35,38,41-45 COF-5 is

formed from the condensation of 2,3,6,7,10,11-hexahydroxytriphenylene (HHTP) and 1,4-

phenylenebis(boronic acid) (PBBA), as shown in Scheme 1. We have previously developed a

kinetic Monte Carlo (KMC) model for 2D COF formation in solution, which makes use of a

combination of experimentally and computationally derived parameters.38 The rate parameters

related to bond formation, bond breakage, and van der Waals interactions involved in the COF-5

growth have been parameterized in our previous work and found to reproduce very well the

experimental production curves.38 Here, in order to fully quantify the nucleation and growth rates,

we carried out two sets of KMC simulations, based on modifications from our original KMC

model.

6

Scheme 1. The two-step growth process of micron-sized crystals obtained by slowly adding

monomers to COF nanoparticle seeds42.

The first set of simulations is designed to evaluate the nucleation rates. Here, the nuclei that have

grown beyond a threshold size of 50 monomers are removed from consideration, thereby

suppressing the impact of growth. The initial system contains 48,200 HHTP molecules and 72,300

PBBA molecules; the volume is varied in order to model different initial monomer concentrations.

These amounts were chosen to minimize any finite-size effects (see Figure S2 in the Supporting

Information, SI) while still allowing for practical simulation times. The simulations are stopped

when the nuclei that are generated correspond to 10% conversion of all monomers; in this way,

the monomer concentrations do not vary significantly from their initial conditions. The nucleation

rates are calculated through a linear fitting of nucleus production with time (additional

computational details can be found in the SI). For each of the initial conditions, 20 simulations

were run in order to reduce the statistical errors.

The second set of simulations focuses on the evaluation of growth rates by preventing additional

nucleation after the appearance of the first nuclei. To model reliably large crystals, the system size

is increased by a factor of ten so as to hold 482,000 HHTP molecules and 723,000 PBBA

molecules. The simulations are completed when the crystal has grown to a size of 100,000

7

monomer units (which corresponds to less than 10% of the total number of monomers). The in-

plane expansion/stacking rates are calculated by considering the evolution of the particle

diameter/height vs. time. Similar to the first set of simulations, in order to obtain different initial

monomer concentrations, the volume is varied while keeping the same number of monomers.

The results of the KMC simulations are shown in Figure 1. They establish the following

fundamental relationships among the nucleation rate (Jnuc), the in-plane (lateral) expansion rate

(vin-plane), and the stacking (vertical expansion) rate (vstacking), with respect to the monomer

concentration (CMonomer = CHHTP = 2/3 CPBBA) –note that, for the sake of simplicity, we neglect the

change in monomer concentration with respect to its initial value:

2

nuc nucleation monomerJ k C= (1)

,in plane growth in plane monomerv k C− −= (2)

,stacking growth stacking monomerv k C= (3)

By considering a series of simulated monomer concentrations, we obtain that the knucleation, kgrowth,in-

plane, and kgrowth,stacking rate coefficients have values of 0.024 L.mol-1.s-1, 1,418 nm.L.mol-1.s-1 and

599 nm.L.mol-1.s-1, respectively.

8

Figure 1. A) KMC-simulated nucleation rate as a function of monomer concentration. B) KMC-

simulated in-plane and stacking growth rates as a function of monomer concentration.

Our earlier KMC simulations showed that nucleation involves bond formation and stacking

between oligomers, which are both second-order in the reactants.38 At steady state, the nucleation

speed is expected to be governed by one of these two steps. Regardless of which step is the limiting

one, a second-order dependence is preserved, as shown by Eq. (1). The first-order dependence of

growth on monomer concentration can be rationalized by considering that, while any growth is

formally a second-order process, given that one reactant is fixed in the crystal, the overall dynamics

become pseudo first-order.

9

Recently, Smith et al. reported an original methodology that allowed them to study COF-5

particles as homogenous colloidal suspensions.41 These COF-5 nanoparticle solutions are much

more suitable for mechanistic investigations than conventional agglomerated particles, since the

heterogeneous processes of precipitation and aggregation are suppressed. Thus, we have measured

the growth rates of the nanocrystalline colloidal suspensions in order to characterize in depth the

mechanism of 2D COF growth.

To probe the rate dependence of COF-5 nucleation or growth, we turned to in situ

Small/Medium/Wide-Angle X-ray Synchrotron Scattering (SAXS/MAXS/WAXS)

measurements. The simultaneous collection of these data sets allowed us to monitor the size and

crystallinity of the particles in tandem, as the reaction proceeded. To probe the nucleation and

growth processes independently, it was necessary to choose conditions under which the processes

were temporally resolved. This was achieved by selecting monomer concentrations under which

one process dominated, as was seen for COF-5 in a recent report by Evans et al.42 To probe the

nucleation behavior of COF-5, we prepared solutions of HHTP and PBBA at concentrations of

over 1 mM, which were recently demonstrated to favor nucleation. After heating these solutions

to 70 °C, we continuously monitored the MAXS pattern during COF-5 nucleation (Figure 2A).

The integrated intensity of the <100> Bragg feature (~0.24 Å-1) as a function of time was extracted

from these data (Figure 2B, inset). Finally, linear fits of signal intensity versus time were extracted

to determine the rate of nucleation under these conditions. We observe a second-order relationship

between the initial monomer concentration and the growth rate of the <100> signal intensity

(Figure 2B). This observation is thus consistent with the results of the KMC simulations described

above and supports a second-order dependence of COF nucleation on monomer concentration.

10

Figure 2. A) MAXS traces as a function of time for selected data of B. B) Rate of intensity of the

<100> Bragg feature as a function of initial monomer concentration (inset: <100> intensity as a

function of time for selected monomer concentrations). C) MAXS traces as a function of time for

selected data of D. D) Rate of increase of intensity of the <100> Bragg feature as a function of

monomer equivalents added (inset: <100> intensity as a function of time for selected equivalents

added).

In order to probe the monomer dependence of growth, we sought to identify conditions under

which nucleation would be suppressed, yet growth would occur at measurable rates. We first

determined an upper concentration limit for growth experiments (0.7 mM) by allowing HHTP and

PBBA to react in the absence of preexisting seed particles. No change in signal was observed,

indicating that nucleation is suppressed at these time scales (Figure S9). It was also critical to

identify concentrations of COF seeds that were observable while sufficiently dilute to allow for

observable growth, which was ultimately optimized to an initial seed concentration. Mixtures of

COF-5 seed particles at a constant particle density but variable free monomer equivalents (moles

HHTP free monomer / moles of HHTP present in COF seeds) below this nucleation threshold were

then prepared and heated to 70 °C (see SI for more details). Additionally, a sample was prepared

11

by diluting the COF-5 seeds in the absence of additional monomers. Once again, the MAXS pattern

was monitored as a function of time (Figure 2C), with extracted <100> intensities (Figure 2D

inset) and rate of this intensity growth (Figure 2D) being determined as described for the

nucleation experiment. In this case, the <100> intensity growth rate is clearly linear with respect

to the monomer equivalents (and therefore the initial monomer concentration) being added. As

expected, under conditions where no additional monomers were added to the diluted COF seeds,

the initial seed particles remained unchanged (Figure S10), which confirms that any enhancement

in signal is not related to processes such as Ostwald ripening. Taken together, these results indicate

that growth is a first-order process with respect to monomer concentration when the particle

density of COF-5 seeds is held constant.

We stress that precise control of nucleation and growth rates is critical to fabricating high-quality

2D COF crystals. Based on the above results, it can be concluded that for COF-5 particles

nucleation is an inherently higher-order process than growth. As we elaborate on in the following

discussion, this difference in dynamics is the origin of the observation that nucleation and growth

can be temporally resolved by controlling the monomer concentration, with nucleation being

suppressed at lower monomer concentrations.42,46

We now turn to a discussion of the relative strengths of nucleation and growth rates. One way to

compare these rates is through monomer consumption. The monomer consumption rates for

nucleation (ωmonomer,nuc) and growth (ωmonomer,growth) of a crystal take on the following forms when

assuming a cylindrical shape and a constant diameter-to-height ratio:

2

, nucNmonomer nuc nuc monomerJ C = (4)

12

2

,monomer growth monomer

dVd C

dt = (5)

where Nnuc is the nucleus size, taken here as 50 monomer units; β, the number of monomer units

per unit volume in the crystal; V, the crystal volume; t, the time; and d, the crystal diameter.

If we consider the monomer consumption rates of a colloidal system consisting of uniform

microcrystals of diameter d, the steady-state nucleation and growth rates are sketched in Figure

3A. Therefore, the second- and first-order dependences of nucleation and growth rates on

monomer concentration mean that there always exists a threshold concentration below which

nucleation is slower than growth (Figure 3A). This analysis thus rationalizes the recent findings

that nucleation is suppressed when adding monomers slowly to the colloidal solution.42,46

Figure 3. A) Comparison of the monomer consumption dynamics of nucleation and growth. B)

Illustration of the critical monomer concentration and reaction time at different desired crystallite

expansion (l). The value of Cnuc,0 is assumed to be 3.3×10-5 mM.

13

Understanding the relative rates of competing processes allows for guided experimental design of

COF crystal growth by selecting a specific, steady-state monomer concentration. Higher monomer

concentrations will speed the growth of existing colloids, yet complicates the suppression of the

nucleation of new particles. Considering the interplay between nucleation and growth processes

are always at play, it is challenging to completely eliminate nucleation. However, with the

objective of suppressing nucleation, we define γ as the number of newly nucleated species divided

by the initial seed crystals (it can be understood as a proportionality factor between the occurrence

of new crystals with respect to the existing ones) during a growth process in which the diameter

of the existing crystals is extended by l. There exists a critical monomer concentration (C*) that

represents the upper theoretical limit for a specific combination of γ and l. By this definition, the

monomer concentration in the reaction needs to be lower than C* to achieve the desired crystal

expansion while suppressing nucleation. Monomer concentrations above C* lead to more frequent

nucleation (increased γ) and smaller average crystal sizes. We find that C* depends on the initial

nuclei concentration (Cnuc,0) and can be expressed as:

,0 ,* nuc growth in plane

nucleation

C kC

lk

−= (6)

The monomer consumption rate ωconsump (from both nucleation and growth) gives information

about the monomer amount that should be added to maintain a constant monomer concentration at

Cmonomer. When assuming the crystals to be cylindrical and under the condition that γ≪1, we find

that ωconsump is expressed as:

( )2, 2

, 0 nuc

3N

4

growth in plane nuclei monomer

consump growth in plane monomer nucleation monomer

k C Ck C t d k C

−= + + (7)

14

where α is the diameter-to-height ratio of the crystal and d0, the initial diameter of the crystals. It

emerges from Eq. (7) that ωconsump is not constant but instead increases quadratically with time.

For practical 2D COF growth, the reaction time is also an important factor to be considered. The

use of too low a monomer concentration should be avoided as it results in undesirably slow growth.

The shortest time for crystal growth is achieved when using Cmonomer = C*. In this case, the time

required for the desired crystal expansion l is calculated to be:

,

2

* 2

, ,0growth in plane

nucleation

growth in plane nuclei

k ll

k C k C

−−

= = (8)

The quadratic dependence of time on desired growth coupled with the inverse dependence on γ

means that growing ever larger crystals by slow monomer addition is likely to face practical

challenges. Nevertheless, this analysis provides a theoretical reference for the monomer

concentration to use during 2D COF fabrication, the amount of monomers to add, and the estimated

time of crystal growth. This theoretical foundation is critical to the rational control over 2D COF

synthetic conditions, and therefore temporal resolution of different microscale processes relevant

to COF growth. Furthermore, the model we have developed for COF growth suggests that

mechanisms other than direct condensation (e.g., formal transimination 47 and monomer exchange

reactions 48) need to be explored to determine whether growth can be favored over nucleation,

which could yield 2D COF single crystals under more practical time scales. In addition, we

anticipate that combinations of strategies will likely be more effective than individual approaches.

15

Conclusions

The combination of kinetic Monte Carlo simulations and in situ X-ray scattering measurements

has allowed us to develop a quantitative analysis of the nucleation and growth rates of 2D COFs,

as represented by COF-5. The nucleation and growth processes have second-order and first-order

dependences on monomer concentration, respectively. These different reaction orders explain why

growth can be suppressed by controlling the speed of monomer addition. Based on this

experimentally validated model, we were able to define a critical monomer concentration C*,

which provides insight into the optimal monomer concentration to use for 2D COF growth via

slow addition of monomers to 2D COF colloidal seeds.

This understanding allows us to express several rules for the optimization of 2D COF synthesis.

First, we find that C* is proportional to the occurrence of new crystals with respect to the existing

ones, while inversely proportional to the desired amount of expansion of the seed. Therefore, lower

monomer concentrations should be selected when targeting fewer and larger crystallite domains.

Secondly, C* is linearly dependent on the number of nuclei in solution. This means that higher

monomer concentrations may be used in solutions where the number of nuclei is also larger.

Therefore, in solutions with more COF nanoparticles, the critical monomer concentration is

increased as well as the rate of monomer consumption directed toward crystal growth. However,

this growth will be shared among the larger number of crystallite seeds, making this strategy

ineffective for increasing the average crystallite domain size more quickly.

The fundamental connection between C* and the competition between nucleation and growth

means that the reaction time has a quadratic dependence on the desired crystal expansion. While

no theoretical upper limit exists in terms of a maximum COF size, the nonlinear dependence on

16

time suggests that growing large crystals can rapidly become impractical. For example, as

illustrated by Eq. (8), growing 1-mm long crystals for a value of γ (the proportionality factor

between the occurrence of new crystals with respect to the existing ones) equal to 0.01, can take

over 1,000,000 years. Thus, growing millimeter-sized 2D COF crystals from solution using a

direct-condensation, seeded-growth approach remains challenging. However, this limitation is

associated with the specific strategy of limiting nucleation only by maintaining a low monomer

concentration. Other methods to limit nucleation, such as chemically increasing the

transesterification rate associated with COF formation, might provide access to high-quality COFs

under different kinetic regimes. New synthetic strategies are called for to achieve these goals.

Finally, we note that 2D COF formation and growth can rely on different mechanisms. Here, we

have used COF-5 as a representative example and the conclusions thus likely apply to the series

of boronate ester-linked COFs. It will be of interest to determine whether the set of empirical

equations developed here can be extended to other types of 2D COFs for which different growth

mechanisms have been proposed. Such mechanistic investigations of chemically distinct systems

are part of our ongoing collaborative efforts.

ASSOCIATED CONTENT

Supporting Information. Additional details on KMC simulations, the calculations of nucleation

and growth rates and in situ XRD measurements. This material is available free of charge via the

Internet at http://pubs.acs.org.

17

AUTHOR INFORMATION

Notes

The authors declare no competing financial interests.

ACKNOWLEDGMENTS

This work was supported by the Army Research Office, under the Multidisciplinary University

Research Initiative (MURI) program, Award No. W911NF-15-1-0447, and under Grant No.

W911NF-17-1-0339. A. M. E. (DGE-1324585), M.J.S. (DGE-1842165), and I.C. (DGE-1842165)

are supported by the National Science Foundation Graduate Research Fellowship. I.C. is partially

supported by the Ryan Fellowship and the International Institute for Nanotechnology. Parts of this

work were performed at the DuPont-Northwestern-Dow Collaborative Access Team (DND-CAT)

located at Sector 5 of the Advanced Photon Source (APS); DND-CAT is supported by

Northwestern University, E.I. DuPont de Nemours & Co., and the Dow Chemical Company. This

research used resources of the Advanced Photon Source and Center for Nanoscale Materials, both

U.S. Department of Energy (DOE) Office of Science User Facilities operated for the DOE Office

of Science by Argonne National Laboratory under Contract No. DE-AC0206CH11357.

REFERENCES

(1) Sakamoto, J.; van Heijst, J.; Lukin, O.; Schlüter, A. D., Two-Dimensional Polymers: Just

a Dream of Synthetic Chemists?, Angew. Chem., Int. Ed. 2009, 48, 1030.

(2) Cote, A. P.; Benin, A. I.; Ockwig, N. W.; O'Keeffe, M.; Matzger, A. J.; Yaghi, O. M.,

Porous, Crystalline, Covalent Organic Frameworks, Science 2005, 310, 1166.

(3) Feng, X.; Ding, X. S.; Jiang, D. L., Covalent Organic Frameworks, Chem. Soc. Rev. 2012,

41, 6010.

(4) Wu, D. C.; Xu, F.; Sun, B.; Fu, R. W.; He, H. K.; Matyjaszewski, K., Design and

Preparation of Porous Polymers, Chem. Rev. 2012, 112, 3959.

18

(5) Colson, J. W.; Dichtel, W. R., Rationally Synthesized Two-Dimensional Polymers, Nature

Chem. 2013, 5, 453.

(6) Ding, S. Y.; Wang, W., Covalent Organic Frameworks (COFs): From Design to

Applications, Chem. Soc. Rev. 2013, 42, 548.

(7) Bertrand, G. H. V.; Michaelis, V. K.; Ong, T.-C.; Griffin, R. G.; Dincă, M., Thiophene-

Based Covalent Organic Frameworks, Proc. Natl. Acad. Sci. 2013, 110, 4923.

(8) Xu, H.; Gao, J.; Jiang, D., Stable, Crystalline, Porous, Covalent Organic Frameworks as a

Platform for Chiral Organocatalysts, Nature Chem. 2015, 7, 905.

(9) Waller, P. J.; Gandara, F.; Yaghi, O. M., Chemistry of Covalent Organic Frameworks, Acc.

Chem. Res. 2015, 48, 3053.

(10) Huang, N.; Wang, P.; Jiang, D., Covalent Organic Frameworks: A Materials Platform for

Structural and Functional Designs, Nat. Rev. Mater. 2016, 1, 16068.

(11) Ascherl, L.; Sick, T.; Margraf, J. T.; Lapidus, S. H.; Calik, M.; Hettstedt, C.; Karaghiosoff,

K.; Döblinger, M.; Clark, T.; Chapman, K. W.; Auras, F.; Bein, T., Molecular Docking Sites

Designed for the Generation of Highly Crystalline Covalent Organic Frameworks, Nature Chem.

2016, 8, 310.

(12) Beuerle, F.; Gole, B., Covalent Organic Frameworks and Cage Compounds: Design and

Applications of Polymeric and Discrete Organic Scaffolds, Angew. Chem., Int. Ed. 2018, 57, 4850.

(13) Lohse, M. S.; Bein, T., Covalent Organic Frameworks: Structures, Synthesis, and

Applications, Adv. Funct. Mater. 2018, 28, 1705553.

(14) Bisbey, R. P.; Dichtel, W. R., Covalent organic frameworks as a platform for

multidimensional polymerization, Acs Central Science 2017, 3, 533.

(15) DeBlase, C. R.; Silberstein, K. E.; Truong, T. T.; Abruna, H. D.; Dichtel, W. R., β-

Ketoenamine-linked covalent organic frameworks capable of pseudocapacitive energy storage, J.

Am. Chem. Soc. 2013, 135, 16821.

(16) Sun, B.; Zhu, C. H.; Liu, Y.; Wang, C.; Wan, L. J.; Wang, D., Oriented Covalent Organic

Framework Film on Graphene for Robust Ambipolar Vertical Organic Field-Effect Transistor,

Chem. Mater. 2017, 29, 4367.

(17) Miao, J.; Xu, Z.; Li, Q.; Bowman, A.; Zhang, S.; Hu, W.; Zhou, Z.; Wang, C., Vertically

Stacked and Self-Encapsulated van der Waals Heterojunction Diodes Using Two-Dimensional

Layered Semiconductors, ACS Nano 2017, 11, 10472.

(18) Dey, K.; Pal, M.; Rout, K. C.; Kunjattu H, S.; Das, A.; Mukherjee, R.; Kharul, U. K.;

Banerjee, R., Selective Molecular Separation by Interfacially Crystallized Covalent Organic

Framework Thin Films, J. Am. Chem. Soc. 2017, 139, 13083.

(19) Medina, D. D.; Sick, T.; Bein, T., Photoactive and Conducting Covalent Organic

Frameworks, Adv. Energy. Mater. 2017, 7, 1700387.

(20) Xiang, Z. H.; Cao, D. P.; Dai, L. M., Well-defined two dimensional covalent organic

polymers: rational design, controlled syntheses, and potential applications, Polym. Chem. 2015, 6,

1896.

(21) Pyles, D. A.; Crowe, J. W.; Baldwin, L. A.; McGrier, P. L., Synthesis of Benzobisoxazole-

Linked Two-Dimensional Covalent Organic Frameworks and Their Carbon Dioxide Capture

Properties, ACS Macro Lett. 2016, 5, 1055.

(22) Xu, Q.; Tao, S.; Jiang, Q.; Jiang, D., Ion Conduction in Polyelectrolyte Covalent Organic

Frameworks, J. Am. Chem. Soc. 2018, 140, 7429.

(23) Zhao, F.; Liu, H.; Mathe, S.; Dong, A.; Zhang, J., Covalent Organic Frameworks: From

Materials Design to Biomedical Application, Nanomaterials 2018, 8, 15.

19

(24) Zhao, W.; Xia, L.; Liu, X., Covalent organic frameworks (COFs): perspectives of

industrialization, CrystEngComm 2018, 20, 1613.

(25) Wan, S.; Guo, J.; Kim, J.; Ihee, H.; Jiang, D., A Belt-Shaped, Blue Luminescent, and

Semiconducting Covalent Organic Framework, Angew. Chem., Int. Ed. 2008, 47, 8826.

(26) Wan, S.; Gándara, F.; Asano, A.; Furukawa, H.; Saeki, A.; Dey, S. K.; Liao, L.; Ambrogio,

M. W.; Botros, Y. Y.; Duan, X.; Seki, S.; Stoddart, J. F.; Yaghi, O. M., Covalent Organic

Frameworks with High Charge Carrier Mobility, Chem. Mater. 2011, 23, 4094.

(27) Duhović, S.; Dincă, M., Synthesis and Electrical Properties of Covalent Organic

Frameworks with Heavy Chalcogens, Chem. Mater. 2015, 27, 5487.

(28) Crowe, J. W.; Baldwin, L. A.; McGrier, P. L., Luminescent Covalent Organic Frameworks

Containing a Homogeneous and Heterogeneous Distribution of Dehydrobenzoannulene Vertex

Units, J. Am. Chem. Soc. 2016, 138, 10120.

(29) Jin, E.; Asada, M.; Xu, Q.; Dalapati, S.; Addicoat, M. A.; Brady, M. A.; Xu, H.; Nakamura,

T.; Heine, T.; Chen, Q.; Jiang, D., Two-Dimensional sp2 Carbon–Conjugated Covalent Organic

Frameworks, Science 2017, 357, 673.

(30) Haldar, S.; Chakraborty, D.; Roy, B.; Banappanavar, G.; Rinku, K.; Mullangi, D.; Hazra,

P.; Kabra, D.; Vaidhyanathan, R., Anthracene-Resorcinol Derived Covalent Organic Framework

as Flexible White Light Emitter, J. Am. Chem. Soc. 2018, 140, 13367.

(31) Albacete, P.; Martínez, J. I.; Li, X.; López-Moreno, A.; Mena-Hernando, S. a.; Platero-

Prats, A. E.; Montoro, C.; Loh, K. P.; Pérez, E. M.; Zamora, F., Layer-Stacking-Driven

Fluorescence in a Two-Dimensional Imine-Linked Covalent Organic Framework, J. Am. Chem.

Soc. 2018, 140, 12922.

(32) Wan, S.; Guo, J.; Kim, J.; Ihee, H.; Jiang, D. L., A Belt-Shaped, Blue Luminescent, and

Semiconducting Covalent Organic Framework, Angew. Chem., Int. Ed. Engl. 2008, 47, 8826.

(33) Sick, T.; Hufnagel, A. G.; Kampmann, J.; Kondofersky, I.; Calik, M.; Rotter, J. M.; Evans,

A.; Döblinger, M.; Herbert, S.; Peters, K.; Böhm, D.; Knochel, P.; Medina, D. D.; Fattakhova-

Rohlfing, D.; Bein, T., Oriented Films of Conjugated 2D Covalent Organic Frameworks as

Photocathodes for Water Splitting, J. Am. Chem. Soc. 2018, 140, 2085.

(34) Gonçalves, R. S. B.; de Oliveira, A. B. V.; Sindra, H. C.; Archanjo, B. S.; Mendoza, M.

E.; Carneiro, L. S. A.; Buarque, C. D.; Esteves, P. M., Heterogeneous Catalysis by Covalent

Organic Frameworks (COF): Pd(OAc)2@COF-300 in Cross-Coupling Reactions, ChemCatChem

2016, 8, 743.

(35) Smith, B. J.; Dichtel, W. R., Mechanistic Studies of Two-Dimensional Covalent Organic

Frameworks Rapidly Polymerized from Initially Homogenous Conditions, J. Am. Chem. Soc.

2014, 136, 8783.

(36) Smith, B. J.; Hwang, N.; Chavez, A. D.; Novotney, J. L.; Dichtel, W. R., Growth Rates and

Water Stability of 2D Boronate Ester Covalent Organic Frameworks, Chem. Commun. 2015, 51,

7532.

(37) Medina, D. D.; Rotter, J. M.; Hu, Y.; Dogru, M.; Werner, V.; Auras, F.; Markiewicz, J. T.;

Knochel, P.; Bein, T., Room Temperature Synthesis of Covalent–Organic Framework Films

through Vapor-Assisted Conversion, J. Am. Chem. Soc. 2015, 137, 1016.

(38) Li, H. Y.; Chavez, A. D.; Li, H. F.; Li, H.; Dichtel, W. R.; Brédas, J.-L., Nucleation and

Growth of Covalent Organic Frameworks from Solution: The Example of COF-5, J. Am. Chem.

Soc. 2017, 139, 16310.

(39) De Yoreo, J. J.; Gilbert, P. U. P. A.; Sommerdijk, N. A. J. M.; Penn, R. L.; Whitelam, S.;

Joester, D.; Zhang, H. Z.; Rimer, J. D.; Navrotsky, A.; Banfield, J. F.; Wallace, A. F.; Michel, F.

20

M.; Meldrum, F. C.; Colfen, H.; Dove, P. M., Crystallization by particle attachment in synthetic,

biogenic, and geologic environments, Science 2015, 349, aaa6760.

(40) Sosso, G. C.; Chen, J.; Cox, S. J.; Fitzner, M.; Pedevilla, P.; Zen, A.; Michaelides, A.,

Crystal nucleation in liquids: Open questions and future challenges in molecular dynamics

simulations, Chem. Rev. 2016, 116, 7078.

(41) Smith, B. J.; Parent, L. R.; Overholts, A. C.; Beaucage, P. A.; Bisbey, R. P.; Chavez, A.

D.; Hwang, N.; Park, C.; Evans, A. M.; Gianneschi, N. C.; Dichtel, W. R., Colloidal Covalent

Organic Frameworks, ACS Central Science 2017, 3, 58.

(42) Evans, A. M.; Parent, L. R.; Flanders, N. C.; Bisbey, R. P.; Vitaku, E.; Kirschner, M. S.;

Schaller, R. D.; Chen, L. X.; Gianneschi, N. C.; Dichtel, W. R., Seeded Growth of Single-Crystal

Two-Dimensional Covalent Organic Frameworks, Science 2018, 361, 52.

(43) Li, H. F.; Li, H. Y.; Dai, Q. Q.; Li, H.; Brédas, J. L., Hydrolytic Stability of Boronate

Ester‐Linked Covalent Organic Frameworks, Adv. Theory Simul. 2018, 1, 1700015.

(44) Li, H. Y.; Brédas, J.-L., Large Out-of-Plane Deformations of Two-Dimensional Covalent

Organic Framework (COF) Sheets, J. Phys. Chem. Lett. 2018, 9, 4215.

(45) Li, H. Y.; Brédas, J.-L., Nanoscrolls Formed from Two-Dimensional Covalent Organic

Frameworks, Chem. Mater. 2019, 31, 3265.

(46) Li, Rebecca L.; Flanders, N. C.; Evans, A. M.; Ji, W.; Castano, I.; Chen, L. X.; Gianneschi,

N. C.; Dichtel, W. R., Controlled growth of imine-linked two-dimensional covalent organic

framework nanoparticles, Chem. Sci. 2019, 10, 3796.

(47) Vitaku, E.; Dichtel, W. R., Synthesis of 2D Imine-Linked Covalent Organic Frameworks

through Formal Transimination Reactions, J. Am. Chem. Soc. 2017.

(48) Daugherty, M. C.; Vitaku, E.; Li, R. L.; Evans, A. M.; Chavez, A. D.; Dichtel, W. R.,

Improved synthesis of β-ketoenamine-linked covalent organic frameworks via monomer exchange

reactions, Chem. Commun. 2019, 55, 2680.

S1

Nucleation-Elongation Dynamics of Two-Dimensional

Covalent Organic Frameworks

Supplementary Information

Haoyuan Lia§, Austin M. Evansb§, Ioannina Castanob, Michael J. Straussb, William R. Dichtelb*

and Jean-Luc Bredasa*

aSchool of Chemistry and Biochemistry, Center for Organic Photonics and Electronics (COPE),

Georgia Institute of Technology, Atlanta, Georgia 30332-0400 bDepartment of Chemistry, Northwestern University, Evanston, IL 60208, USA.

Corresponding Authors

*[email protected]

*[email protected]

Table of Contents

I. Simulation Results ............................................................................ S2

II. Materials and Instrumentation ...................................................... S7

III. Synthesis of COF-5 Seeds ............................................................. S8

IV. Experimental in situ Scattering Details ....................................... S9

V. X-Ray Scattering Control Experiments ...................................... S12

VI. In Situ X-Ray Scattering Nucleation Results ............................ S14

VII. In Situ X-Ray Scattering Growth Results ................................ S20

S2

I. Simulation Results

The parameters used in the KMC simulations have been documented in our previous work.1 In the

present KMC simulations, nuclei have been identified by setting a size threshold of 50 monomer

units, which we choose to be slightly larger than the value of 40 monomer units used in our

previous study 1 in order to ensure better identification. Figure S1 shows a typical nucleus

production curve. For each simulation, a linear fitting is performed. The nucleation rate

corresponds to the fitted slope.

Figure S1. Representative Linear Fitting Used in the KMC Simulations: Illustration of the

calculation of the nucleation rates.

S3

For nucleation simulations, the total number of monomers is 120,500. We have checked that the

error coming from the finite sizes of the simulation system is not significant, as shown in Figure

S2.

Figure S2. Simulation Results for Different System Sizes in the KMC Simulations. N is the

total number of monomers. The concentrations of monomers are the same across all simulations.

The lateral expansion is found to be approximately linear with time. Figures S3 and S4 show a

typical COF expansion in the course of the simulated growth and the top view of the final crystal,

respectively. Linear fittings are performed on the crystal expansion results, see Figure S5. The

lateral expansion rate thus corresponds to the fitted slope.

S4

Figure S3. Example of KMC Growth Result. Illustration of the evolution of average diameter

and height as a function of time.

S5

Figure S4. COF-5 particle simulation. Illustration of the final structure of the COF-5 crystal (top

view and side view) from the KMC simulation.

S6

Figure S5. Fitting of the Lateral Expansion Data. Lateral expansion as a function of time and

linear fittings.

S7

II. Materials and Instrumentation

Materials

Acetonitrile, 1-4 Dioxane, and Mesitylene were purchased from Fisher Scientific.

Phenylbisboronic acid (PBBA) was purchased from Sigma-Aldrich and hexahydroxytriphenylene

(HHTP) was purchased from TCI America. All chemicals were used without further purification.

Instrumentation

Sonication. Sonication was performed with a Branson 3510 ultrasonic cleaner with a power

output of 100W and a frequency of 42 kHz.

In Situ X-ray Scattering: Small/Medium/Wide-Angle X-ray Scattering (SAXS/MAXS/WAXS)

data were collected simultaneously at sector 5-ID-D of the Advanced Photon Source at Argonne

National Laboratory. A beam energy of 17.0 KeV was used for these experiments. Patterns were

collected with single 10 s frames on a series of 3 Pilatus 2D detectors which were then radially

integrated. All samples were conducted in 1.5 mm borosilicate capillaries with a wall thickness of

0.01 mm available from Charles Supper Scientific. Patterns were assigned to be time = 0 seconds

when the heater reached its steady state temperature of 70 °C (approximately 45 seconds). Patterns

were baselined and then background subtracted from the starting time pattern to produce corrected

patterns. The <100> Bragg feature of COF-5 (~0.24 Å-1) was then integrated and plotted against

the time of each pattern. Finally, the rate of increase in <100> signal intensity was determined by

fitting the slope of the respective <100> intensity versus time for each experiment.

S8

III. Synthesis of COF-5 Seeds

Figure S6. Synthesis Procedure of COF-5 Colloidal Seeds.

Hexahydroxytriphenylene (HHTP) and phenylbisboronicacid (PBBA) were added to a 20 mL

mixture of CH3CN:dioxane:mesitylene (80:16:4 vol%) at a concentration of 2 mM HHTP and 3

mM PBBA. This solution was sonicated until full dissolution of the monomers was observed. This

solution was then heated in a scintillation vial overnight at 70 °C after which time the solution

became opaque indicating the successful synthesis of COF-5 seed particles. These seeds were then

used as prepared for COF growth experiments detailed below. This method is explained in further

detail in Ref. 2

S9

IV. Experimental in situ Scattering Details

Figure S7. Nucleation Experiment. Schematic of how the nucleation experiments were

conducted.

Nucleation Experiment: Hexahydroxytriphenylene (HHTP) and phenylbisboronicacid (PBBA)

were added to a 20 mL mixture of CH3CN:dioxane:mesitylene (80:16:4 vol%) in a 1:1.5 molar

ratio at a concentration of double that specified in the experiment. For example, if the nucleation

experiment was conducted at 2 mM, an HHTP solution of 4 mM and a PBBA solution of 6 mM

would have been made independently. This is to prevent reaction before the mixing of the two

monomers. These solutions were sonicated until full solvation of the monomers. The two monomer

solutions were then mixed in a dram vial in a 1:1 volume ratio (final volume 1 mL) and were then

diluted as specified to produce the specified monomer concentration. These solutions were then

quickly shaken to ensure mixing. Approximately 100 µL of these solutions were then added to a

borosilicate capillary and sealed using capillary sealing putty. These capillaries were immediately

installed onto a multicapillary heating stage available at Sector 5-ID-D of the Advanced Photon

S10

Source (APS) Argonne National Lab (ANL). This stage was then heated to 70 °C (< 1 min heating

time) and allowed to react while continuous SAXS/MAXS/WAXS patterns were taken.

Figure S8. Growth Experiment. Schematic of how the growth experiments were conducted.

Growth Experiment: Hexahydroxytriphenylene (HHTP) and phenylbisboronicacid (PBBA)

were added to a separate 20 mL mixture of CH3CN:dioxane:mesitylene (80:16:4 vol%) at a

concentration of 1 mM HHTP and 1.5 mM PBBA. These solutions were sonicated until full

solvation of the monomers. The two monomer solutions were then mixed in a dram vial in a 1:1

volume ratio (final volume 2 mL) and quickly shaken. Dilutions were then performed in dram vials

to produce different concentrations of monomer solutions with the same volume. The highest

monomer concentration (0.48 mM) was chosen to temporally resolve nucleation and growth. The

same amount of COF-5 seeds (COF-5 seeds diluted by a factor of 20, resulting in particle solution

containing 0.1 mM HHTP and 0.15 mM PBBA) were then added to all reactions to keep the density

of seed particles identical. To ensure that Ostwald ripening or depolymerization were not relevant

processes under these conditions, we also diluted our COF-5 seed particles and performed the

S11

experiment without the presence of additional monomers. Precise amounts of the dilution

experiment are recorded below.

Table S1. Dilution Calculations. Calculations of how dilutions were conducted for COF-5

growth experiments.

Approximately 100 µL of these solutions were then added to a borosilicate capillary and sealed

using capillary sealing putty. These capillaries were immediately installed onto a multicapillary

heating stage available at Sector 5-ID-D of the Advanced Photon Source (APS) Argonne National

Lab (ANL). This stage was then heated to 70 °C and allowed to react while continuous

SAXS/MAXS/WAXS patterns were taken.

Seed Stock

(mL)

Blank Solvent

(mL)

Monomer Amount

(mL)

Resultant Monomer

Concentration (mM)

0.1 0 1.9 0.48

0.1 0.4 1.5 0.38

0.1 0.6 1.3 0.33

0.1 0.8 1.1 0.28

0.1 1 0.9 0.23

0.1 1.2 0.7 0.18

0.1 1.9 0 0

S12

V. X-Ray Scattering Control Experiments

Figure S9. Nucleation Control. In situ X-ray scattering of 0.7 mM HHTP 1.1 mM PBBA solution

over the course of 1 hour.

The background subtracted MAXS data shown in the bottom right corner demonstrates that under

conditions where monomer concentration is sufficiently small results in no <100> diffraction

feature and thus no substantial COF nucleation on the time-scales studied here. The background

subtracted SAXS data shown in the bottom left corner shows that species of small sizes are formed

over the course of this reaction. This validates previous understanding of COF nucleation where

small oligomers are formed but never “nucleate” unless they reach a critical concentration.

S13

Figure S10. Growth Control. In situ X-ray scattering of COF-5 seeds heated in the presence of

no additional monomer.

The MAXS data shown in the top right corner demonstrates that under conditions where no

additional monomer is added no discernable difference in the <100> diffraction feature is noted.

The background subtracted SAXS data shown in the bottom left corner shows that there is no

discernable size change of the existing nucleation after the course of one hour. These experiments

suggest that the increase in intensity seen under monomer addition conditions are due to growth

and not processes such as Ostwald ripening. This is consistent with observations made previously.2

S14

VI. In Situ X-Ray Scattering Nucleation Results

All experiments were conducted over the course of approximately one hour. Conditions of each

experiment are given in their respective figure headings. In these figures [Monomer] = [HHTP] =

2/3*[PBBA] is the concentration of monomers present at the beginning of the experiment.

Figure S11. Nucleation Experiment [Monomer] = 0.9 mM (measurement 1).

S15

Figure S12. Nucleation Experiment [Monomer] = 0.9 mM (measurement 2).

Figure S13. Nucleation Experiment [Monomer] = 1.1 mM (measurement 1).

S16

Figure S14. Nucleation Experiment [Monomer] = 1.1 mM (measurement 2).

Figure S15. Nucleation Experiment [Monomer] = 1.3 mM (measurement 1).

S17

Figure S16. Nucleation Experiment [Monomer] = 1.3 mM (measurement 2).

Figure S17. Nucleation Experiment [Monomer] = 1.5 mM.

S18

Figure S18. Nucleation Experiment [Monomer] = 1.8 mM.

Figure S19. Nucleation Experiment [Monomer] = 2.3 mM.

S19

Figure S20. Nucleation Experiment [Monomer] = 3.0 mM.

S20

VII. In Situ X-Ray Scattering Growth Results

All experiments were conducted over the course of approximately one hour. Conditions of each

experiment are given in their respective figure headings. In these figures [Monomer] = [HHTP] =

2/3*[PBBA] is related to the amount of monomer added to a consistent amount and concentration

of preexisting COF-5 seeds.

Figure S21. Growth Experiment [Monomer] = 0.18 mM.

S21

Figure S22. Growth Experiment [Monomer] = 0.23 mM.

Figure S23. Growth Experiment [Monomer] = 0.28 mM.

S22

Figure S24. Growth Experiment [Monomer] = 0.33 mM.

Figure S25. Growth Experiment [Monomer] = 0.38 mM.

S23

Figure S26. Growth Experiment [Monomer] = 0.48 mM.

References

(1) Li, H. Y.; Chavez, A. D.; Li, H. F.; Li, H.; Dichtel, W. R.; Brédas, J.-L., Nucleation and

Growth of Covalent Organic Frameworks from Solution: The Example of COF-5, J. Am. Chem.

Soc. 2017, 139, 16310.

(2) Evans, A. M.; Parent, L. R.; Flanders, N. C.; Bisbey, R. P.; Vitaku, E.; Kirschner, M. S.;

Schaller, R. D.; Chen, L. X.; Gianneschi, N. C.; Dichtel, W. R., Seeded Growth of Single-Crystal

Two-Dimensional Covalent Organic Frameworks, Science 2018, 361, 52.