Covalent Organic Framework For Efficient Two-Photon Absorption

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Article

Covalent Organic Framework for Efficient


Two-Photon Absorption
Liang Zhang, Yi Zhou, Mei
Jia, ..., Zhihong Liu, Jun Cheng,
Hexiang Deng
hdeng@whu.edu.cn

HIGHLIGHTS
Collectively alignment of
chromophores in COF crystal
lattice

Serrated packing between the


layers

2D imine-linked COF single


crystals with micrometer size

High two-photon absorption


efficiency

Crystallizing molecules to construct covalent organic frameworks (COFs) is an


effective way to achieve precise spatial arrangement, whereby their transition
dipoles were collectively aligned toward the same direction. The linkage and
crystallinity allow for precise control of the orientation of molecular chromophores
within each two-dimensional (2D) layer, while the unusual serrated packing
between the layers further enhances the overall transition dipoles. These features
of 2D COFs, combined with the tunable functional groups, offer excellent two-
photon absorption properties.

Zhang et al., Matter 2, 1049–1063


April 1, 2020 ª 2020 Elsevier Inc.
https://doi.org/10.1016/j.matt.2020.01.019
Article
Covalent Organic Framework
for Efficient Two-Photon Absorption
Liang Zhang,1,2,10 Yi Zhou,1,7,10 Mei Jia,3,10 Yiwen He,1 Wei Hu,4 Qi Liu,1 Jing Li,5 Xiaohui Xu,1
Chao Wang,1 Anna Carlsson,6 Sorin Lazar,6 Arno Meingast,6 Yanhang Ma,7 Jun Xu,8 Wen Wen,9
Zhihong Liu,4 Jun Cheng,3 and Hexiang Deng1,2,11,*

SUMMARY Progress and Potential


We report the collective alignment of molecular chromophores in the backbone Two-photon absorption (2PA)
of covalent organic frameworks (COFs) as a new chemical approach to construct materials are useful in optical
efficient two-photon absorption (2PA) materials. Six imine COFs were synthe- devices as well as biomedical
sized from aldehyde-terminated triarylamine linkers and p-phenylenediamine. imaging and treatment. Molecular
Attainment of these COFs in single-crystal form revealed the underlying layered chromophores are advantageous
structures and precise spatial arrangements of the constituent chromophores in tuning the 2PA properties by
including an unusual serrated packing between the layers. Coherent orientation different functional groups;
of chromophores within each layer made possible collective alignment of their however, it is challenging to
transition dipoles, while the serrated packing between the layers maximized precisely control the spatial
the crystal anisotropy. As a result, the full potential of molecular chromophores arrangement of the
for 2PA was developed. Up to 110-fold enhancement of two-photon action chromophores and their
cross-section (2PACS) was achieved for the COFs relative to their constituents, underlying molecular dipoles. In
with one member exhibiting a high 2PACS value of 8,756 GM/chromophore- this work, we present the linking of
standing among the best molecule-based 2PA materials. molecular chromophores to form
COF crystals as a new way to
achieve high 2PA absorption
INTRODUCTION
efficiency. The breakthroughs in
Macroscopic optical properties of organic materials are determined by the molecu-
both crystallization and structure
lar dipoles of the constituents, where their spatial arrangement in the structure is crit-
characterizations made it possible
ical.1–3 The challenge is that such arrangements are highly sensitive to chemical func-
to study the correlation between
tionality, frustrating attempts to control connectivity, geometry, spatial orientation,
structural features of COFs and
and packing of the molecular constituents. Two-photon absorption (2PA) is a typical
the 2PA property, valuable for the
non-linear optical phenomenon, important for biomedical treatment, which is cho-
design of molecule-based 2PA
sen for the illustration of the power to control spatial arrangement of molecules.4,5
materials. The critical factors are:
There are two important considerations in achieving desirable 2PA performance in
(1) collective alignment of
material design. Firstly, molecules must be intimately connected to maximize inter-
chromophores; (2) serrated
molecular communication between their electronic structures6–9 and secondly, they
packing between two-
must be crystalline to allow for the formation of large coherent domains, across
dimensional layers; (3) size of
which their transition dipoles can be magnified.10–12 These requirements have
crystal domains; and (4) types of
been realized; however, they are yet to meet together in the current design of
functional groups. The first and
organic optical materials, polymers, and discrete molecular crystals. In this report,
second points are unique merits
we show that covalent organic frameworks (COFs),13,14 in which molecules are linked
from COFs and also the key to
covalently and periodically across the crystal lattice, can satisfy both requirements.
their high 2PA performance.
The linkage and crystallinity, respectively allow for precise control of the spatial
arrangement of molecules to give the needed connectivity, and the collective align-
ment of these molecules in the COF structures, thus maximizing the overall transition
molecular dipoles (Figure 1). Another advantage unique to COFs is the availability
of sufficient pore space to accommodate functional groups on the building
blocks without interfering with their linkage or alignment.15–22 Here, we linked six

Matter 2, 1049–1063, April 1, 2020 ª 2020 Elsevier Inc. 1049


Figure 1. Illustration of the Difference in Spatial Arrangement of Molecular Building Blocks and
Their Underlying Transition Dipoles between Polymer and 2D COFs
In polymers the monomers are randomly oriented, so their transition dipoles are not effectively
accumulated. In contrast, the precise arrangement of molecules in COFs leads to collective
alignment of their transition dipoles, achieving desirable anisotropy and maximization in two-
photon absorption.

tetra-topic chromophores (L-H, L-Ph, L-2MeO, L-2F, L-4F, and L-BT; Figure 2) into six
two-dimensional (2D) COFs, COF-601, COF-602, COF-603, COF-604, COF-605,
and COF-606, which were obtained in single-crystal form. Examination of 2PA prop- 1Key Laboratory of Biomedical Polymers -
erties of these COFs revealed that they offered 16- to 110-fold increase in their two- Ministry of Education, College of Chemistry and
photon action cross-section (2PACS) values in comparison with their corresponding Molecular Sciences, Wuhan University, Wuhan
430072, China
monomers (Figure 6C). COF-606-1 mm, with micrometer-sized crystals and serrated
2The Institute for Advanced Studies, Wuhan
interlayer packing, exhibited the highest 2PACS value (8,756 Göppert-Mayer units University, Wuhan 430072, China
[GM] per chromophore) of any organic material, demonstrating the power of this 3Department of Chemistry, College of Chemistry
COF approach. and Chemical Engineering, Xiamen University,
Xiamen 361005, P.R. China
4Key Laboratory of Analytical Chemistry for
The 2D COFs described here are distinguished from other COFs in that the packing
Biology and Medicine (Ministry of Education),
of the molecular constituents between the layers is resolved experimentally for the College of Chemistry and Molecular Sciences,
first time and found to be serrated, which is distinctly different from the eclipsed Wuhan University, Wuhan 430072, China
5Key Laboratory of Photochemical Conversion
and staggered scenarios commonly applied in the structure simulations and refine-
and Optoelectronic Materials, Technical Institute
ments of COFs.23 The resolution of our COF crystals allowed us to uncover and of Physics and Chemistry, Chinese Academy of
assess the qualitative and quantitative impact of their structural features (geometry, Sciences, Beijing 100190, China
spatial orientation, and packing) on their non-linear optical performance as exempli- 6Thermo Fisher Scientific, Materials & Structural
fied by 2PA. Such information is difficult to obtain from polymers and molecular crys- Analysis, Achtseweg Noord 5, 5651 GG
Eindhoven, the Netherlands
tals, but we find that it is extremely valuable for understanding the optical behavior. 7Schoolof Physical Science and Technology,
These structural features provide an extra handle on the design of organic materials ShanghaiTech University, Shanghai 201210,
that goes beyond the traditional tuning of functionality. China
8NationalCenter for Magnetic Resonance in
Wuhan, State Key Laboratory of Magnetic
RESULTS AND DISCUSSION Resonance and Atomic and Molecular Physics,
Wuhan Institute of Physics and Mathematics,
Designed Synthesis of 2D COFs and Characterizations Chinese Academy of Sciences, Wuhan 430071,
The typical synthesis of these COFs was achieved by solvothermal reaction of alde- China
9Shanghai Synchrotron Radiation Facility,
hyde molecular chromophores with aniline building blocks, generating imine link-
Shanghai Advanced Research Institute, Chinese
ages (Figure S3). Particles of micrometer sizes were obtained by programming the Academy of Sciences, Pudong New Area,
heating profile during synthesis to give accurate control during the nucleation pro- Shanghai 201204, P.R. of China
10These authors contributed equally
cess (Section S2 of Supplemental Information). The connectivity between the molec-
11Lead Contact
ular building blocks in these COFs was investigated by a series of spectroscopic
studies. The successful formation of imine linkage was evidenced by the appearance *Correspondence: hdeng@whu.edu.cn
of a 1,612 cm1 peak, the telescopic vibration of imine bond (nC=N), in the Fourier https://doi.org/10.1016/j.matt.2020.01.019

1050 Matter 2, 1049–1063, April 1, 2020


Matter 2, 1049–1063, April 1, 2020 1051
Figure 2. Linkage of Molecular Monomers Containing Triarylamine Functionalities to Construct COF Crystals
Six tetra-topic chromophores, 4,4 0 ,4 00 ,4%-([1,1 0 -biphenyl]-4,4 0 -diylbis(azane-triyl))tetrabenzaldehyde (L-H) and its derivatives with various functional
groups, phenyl (L-Ph), methoxy (L-2MeO), difluoro (L-2F), tetrafluoro (L-4F), and benzothiadiazole (L-BT), were used for the construction of six COFs,
COF-601, COF-602, COF-603, COF-604, COF-605, and COF-606, respectively. R1 to R4 represent the functional groups in the corresponding linker.

transform infrared spectrum of these COFs (Figures S4–S9). X-ray photoelectron


spectroscopy was also performed, whereby a 398.5-eV peak in the N 1s spectrum
demonstrated the presence of the imine bond (Figures S16–S22). The conversion
of the aldehyde was further confirmed by the disappearance of the 191.5-ppm
peak in the 13C cross-polarization and magic-angle spinning nuclear magnetic reso-
nance (MAS NMR) spectrum of COF-606, and the appearance of the 159-ppm peak,
typical for the imine bond (Figure S10). Similar phenomena were observed in all
other COFs (Figures S11–S15). The uniform size and purity of these COF crystals
was revealed by their corresponding scanning electron microscopy (SEM) images
and dynamic light scattering (DLS) of their aqueous solution (Figures S50–S56 and
S96–S98). These COFs exhibited excellent thermal stability, as no obvious weight
loss was observed before 500 C in the thermal gravimetric analysis (Figures S41–
S47). The permanent porosity of these COFs was studied by N2 isotherm at 77 K,
based on which the Brunauer-Emmett-Teller surface areas were derived, ranging
from 470 to 1,030 m2 g1 (Figures S23–S40).

Characterization of COF Single Crystals


The crystallinity of these COFs was confirmed by the clearly resolvable peaks in their
powder X-ray diffraction patterns (PXRD) collected using both laboratory and syn-
chrotron X-ray sources (Sections S7 and S10 of Supplemental Information). High-res-
olution SEM imaging revealed the layer morphology of these 2D crystals and their
micrometer sizes (Figure 3D and Section S6 of Supplemental Information). To inves-
tigate the coherent domains in each particles in COF samples, we used transmission
electron microscopy (TEM) to examine the long-range ordering of molecules within
the layers, where clear lattice fringes were observed in the crystals of COF-606 (Fig-
ure 4A). Fourier diffractograms were obtained from different regions of a single par-
ticle to give identical orientation of the spots with similar intensity. This, combined
with the continuous fringes extending throughout the particle, revealed that the
entire crystal is one single coherent domain (Figures 4B and 4C).

Structure Determination of These COFs


The knowledge on geometry, spatial orientation, and packing of the molecular con-
stituents in the COFs was attained by detailed structure analysis of their single crys-
tals. Although various types of packing between layers in 2D COFs have been pro-
posed and simulated, so far there is no experimental evidence to unambiguously
identify the packing mode for any 2D COFs. Usually the information along the c
axis (reflecting the distance and packing between layers) was buried under the
wide peaks in PXRD patterns of 2D COFs due to low crystallinity. Recently, three-
dimensional (3D) and 2D COFs with large coherent domains, and even in single-crys-
tal form, were successfully synthesized.24–28 However, the preferred orientation of
the 2D COF crystals made it challenging to align the crystal along the appropriate
zone axis, thus compromising the determination of their packing mode.26–28 In
this study, the sizes of the COF crystals are beyond 1 mm in all three dimensions,
ideally suited for electron diffraction (ED) analysis (Figures 3D and 4A). Sharp diffrac-
tion peaks were observed in the ED analysis of COF-606-1 mm crystals with interpret-
able resolution of 2.9 Å, demonstrating their underlying high crystallinity (Figure 4D).
3D electron diffraction tomography (3D-EDT) data were collected on these COFs,
which allowed for the observation of diffraction patterns along a* and b* of the

1052 Matter 2, 1049–1063, April 1, 2020


Matter 2, 1049–1063, April 1, 2020 1053
Figure 3. COF-606 Obtained in Single-Crystal Form
(A) Structure details of 2D layered COF-606 crystals with serrated packing between the layers.
(B) Presence of C2 axis between benzothiadiazole building blocks and offset between propeller triarylamine units in different layers.
(C) PXRD patterns of COF-606 samples. Red circles, black line, green line, and purple bars represent the experimental PXRD patterns, the calculated
patterns based on Pawley refinement, the difference between experimental data and the calculated patterns, and the Bragg positions, respectively.
(D) SEM image of COF-606 with micrometer size. Scale bar, 500 nm.

crystal (Figures 4E and 4F, respectively), providing the first experimental evidence
for the identification of the packing mode (a*, b*, and c* represent the lattice vectors
in the reciprocal space). The measurements of the distance along a* and b* revealed
a c value of roughly 8 Å. The accurate lattice was determined by Pawley refinement
using PXRD (a = 34.22 Å, b = 46.73 Å, c = 7.93 Å, V = 12,633 Å3, and a = g = 90 ,
b = 95 ), with acceptable residue (Rwp = 5.0%, Rp = 2.7%). This c value was distinc-
tively different from the 3–4 Å commonly assumed in the structure refinement of 2D
COFs with eclipsed packing between layers, unveiling the presence of ABAB pack-
ing (Figure 3B).

Quantifying the Offset between Layers


The quantitative offset between the propeller-shaped triarylamine units at adjacent
layers of COF-606 crystals was determined from both XRD and scanning transmis-
sion electron microscopy (STEM). Prior to the structure confirmation using PXRD pat-
terns, the space group was determined by 3D-EDT analysis. There were three
possible space groups, C2, Cm, and C2/m, based on the extinction conditions of
the diffraction.29 The detailed analysis of reflection intensities in the PXRD patterns
ruled out the possibility of C2/m space group (Figures S59 and S60). The underlying
crystal symmetry was also reflected in the macroscopic morphology of COF crystals
in their SEM images (Figures 3D and S61). The absence of mirror symmetry normal to
the b axis combined with the conformation of the benzothiadiazole motifs relative to
the adjacent benzene rings in the crystal structure of the monomer pointed to C2 as
the most probable space group (Figures 3D and S63). According to the C2 space
group, the shift between triarylamine units in adjacent layers took place along the
a axis. To quantify the offset between layers, we plotted the intensity ratio between
130 and 040 reflections in the PXRD patterns against the possible shift distances (0 to
a, with 1/8 a as interval, where 0 and 1/2 a represent the eclipsed and staggered
packing mode, respectively; Figure S59). The experimental ratio was obtained
based on PXRD patterns of three parallel samples. Compassion of this ratio to the
plot allowed us to narrow the possibility to either 1/8 a or 1/4 a (Figure 5B). The rela-
tive intensity of other reflections in the PXRD pattern, for example 040 and 310, ruled
out the possibility of 1/4 a, thus the exact packing mode between the layers in COF-
606 was identified as serrated with 1/8 a offset along the a axis. Furthermore, this
serrated packing was visualized clearly by STEM imaging using an integrated differ-
ential phase contrast (iDPC) method.30 In contrast to conventional TEM or annular
dark-field imaging, this DPC method allowed for direct measurement of the phase
transmission function. It also promoted the data-collection efficiency, thus inducing
less damage to the COF samples. The offset distance between layers was 0.4 nm as
observed from the iDPC image along the [001] zone axis (Figure 3G), consistent with
the value obtained from the XRD analysis (1/8 a = 0.42 nm).

Two-Photon Action Cross-Section Measurements


As a typical non-linear optical phenomenon, 2PA behavior of a material can be iden-
tified by the power-dependent photoluminescence test. The values of the slopes in
the corresponding plots of 2D COFs in this study were found to be close to 2 (Fig-
ure S110), unambiguously confirming the nature of 2PA behavior in these COFs.
The 2PA efficiency of a material is assessed from the 2PACS, which is a product of

1054 Matter 2, 1049–1063, April 1, 2020


Figure 4. Characterizations of COF Single Crystal by High-Resolution TEM and 3D-EDT
(A) Visualization of lattice fringes of one COF-606-1 mm particle under TEM. Scale bar, 100 nm.
(B) Zoomed-in TEM images of four distinct regions of COF-606 single crystals shown in (A) (green, teal, magenta, and red, respectively). Scale bar, 10 nm.
(C) Fast Fourier transform of these four TEM images giving identical patterns, demonstrating the existence of one large single coherent domain in this
COF crystal. Scale bar, 0.25 nm 1 .
(D) Selected-area electron diffraction pattern of COF-606 along c axis. Scale bar, 1 nm 1 .
(E and F) Projection view along a* (E) and b* (F) from the reconstructed reciprocal lattice. Scale bar, 1 nm 1 .
(G) iDPC image of COF-606 along c axis measured under the STEM model. The intensity of the signal reflects the phase-transition function of COF-606.
Scale bar, 3 nm.

Matter 2, 1049–1063, April 1, 2020 1055


1056 Matter 2, 1049–1063, April 1, 2020
Figure 5. Illustration of COF-606 with Various Packing between Layers
(A) Display of possible crystal structures with various offset along a axis (0–7/8 a, with 1/8 a as
interval, where 0 and 1/2 a represent the eclipsed and staggered packing mode, respectively).
(B) Plot of 130/040 peak intensity ratio of different offset packing models with that of the
experimental data (n = 4).

2PA section value (d) and fluorescence quantum yield (h).7,31–34 The impact of
different functional groups on their 2PA performance was investigated by increasing
the strength of donors and acceptors and the length of the conjugation path. The
molecular chromophores were designed to have a central di-topic motif linked by
two electron-donating triarylamine motifs at the two ends (Figure 2). Functional
groups of different electronic properties were introduced in the central motif of
the chromophore to optimize the 2PA performance. Three classes of chromophores
with different electronic configurations were used for the construction of the corre-
sponding COFs (Figure S3). The direct linkage of triarylamine motif and the inser-
tions of neutral phenyl units give the D-p-D configuration (COF-601 and -602).
Methyl substitution at the central phenyl unit functions as the electron-donating
motif to give a D-p-D-p-D configuration (COF-603). Difluoro, tetrafluoro, and ben-
zothiadiazole substitutions function as the electron-withdrawing motifs to give a
D-p-A-p-D configuration (COF-604, COF-605, and COF-606). The impact of
different electronic configurations was investigated by comparing the 2PA proper-
ties of these COFs with identical underlying topology and similar particle size
(around 500 nm). The 2PACS values of these COFs range from 437 to 5,524 GM
per chromophore (1 GM = 1050 cm4$s$photon1) (Figure 6C). As the distance be-
tween the electron-donating triarylamine motif increases, the 2PACS value increases
from 437 to 1,000 GM per chromophore, due to the enlarged dipole moment, which
is consistent with theoretical studies of molecular chromophores. In the comparison
of COFs with the same chromophore size, a general order was observed, D-p-D-p-
D < D-p-D < D-p-A-p-D. This tendency is in good accordance with the 2PA perfor-
mance of their corresponding building blocks, which can be explained by the influ-
ence of electric dipole on the molecular chromophores, where the transition dipole
moment is largely enhanced by the sandwiching electron-withdrawing functional
groups in between the electron-donating groups (D-p-A-p-D). Among the molecu-
lar monomers, L-BT, with D-p-A-p-D configuration, exhibited the highest 2PACS
value, 187 GM. After linking into COF with 500-nm particle sizes, the 2PACS value
of this COF-606-500 nm was boosted to 5,524 GM/chromophore, 30 times that of
the monomer (Figure 6C). Similarly, drastic promotion in 2PACS values was
observed in other COFs in this study, with 16- to 110-fold increases in comparison
with their corresponding monomers, demonstrating the power of connectivity and
crystallinity (Figures 6C and S109–S125). In contrast, only up to 10-fold increase
was observed by polymerization,7 and the 2PACS value of the resulting polymers
was far exceeded by that of the COFs in this study (Table S10).

The impact from the coherent domain, which reflects both the orientation and
spatial arrangement of chromophores in the structure, was investigated by
comparing COF crystals with different particle sizes but of the same type. COF-
606 with different particle sizes—COF-606-100 nm, COF-606-500 nm, and COF-
606-1 mm—were synthesized with their corresponding non-crystalline polymer,
BT-amorphous-500 nm, as control. The particle size and morphology of these sam-
ples were revealed by both DLS and SEM (Section S10 of Supplemental Information).
Their crystallinity was confirmed by PXRD, while no diffraction was observed for the
control polymer (Figure S98). As the particle size of COF increased, sharper peaks
were observed in the PXRD patterns, and higher uptake in the N2 isotherms
measured at 77 K (Figure S100). The one-photon fluorescence intensity was also

Matter 2, 1049–1063, April 1, 2020 1057


Figure 6. Contributions from Structural Features, Including the Geometry, Spatial Orientation, and Packing, to the 2PA Performances
(A) Structural illustration of monomer and models with different geometry and packing of monomers displayed with the calculated transition dipoles by
TD-DFT.
(B) HOMO and LUMO of 1-pore model with appropriate alignment of monomers.
(C) 2PACS value of molecular chromophores with various functional groups and their corresponding COFs (n = 3).
(D) 2PACS value of COF-606 with different coherent domain size (n = 3).
(E) Calculated transition dipole square of monomer and models with different geometry and packing of monomers.

1058 Matter 2, 1049–1063, April 1, 2020


increased accordingly (Figure S107), while the band gap decreased from 2.09 to
2.04 eV from COF-606-100 nm to COF-606-1 mm, respectively (Sections S8 and
S10 of Supplemental Information). More importantly, as the size of single crystals
became larger, the size of the coherent domain expanded drastically and in turn pro-
moted the 2PA performance (Figure 6D). A 3-fold improvement in 2PACS values was
observed as the particle size increased from 100 nm to 500 nm, while that of COF-
606-500 nm was 6-fold that of the amorphous sample, BT-amorphous-500 nm,
with identical particle size. This clearly outlined the advantage of crystallization in
achieving better alignment of the molecules. As the particle size further increased
to 1 mm, the 2PACS value reached 8,756 GM/chromophore, representing a new
benchmark for organic materials (Figure 6D and Section S11 of Supplemental
Information).

Time-Dependent Density Function Theory Simulations of Various Models


The packing between the layers was also found to be critical for the 2PA perfor-
mance of these COFs. Time-dependent density functional theory (TD-DFT) calcula-
tions were carried out on a series of models with different connectivity, coherent
domain size, and packing between chromophores (monomer, 1-pore-amorphous,
1-pore, 4-pore, 1-pore-eclipsed, 1-pore-serrated, and 1-pore-staggered), to assess
the contributions from their transition dipole moment (Figure 6A and Section S9 of
Supplemental Information). Comparison of the transition dipoles of monomer,
1-pore-amorphous, 1-pore, and 4-pore revealed the same trend as observed in
the study of COFs with different coherent domain sizes. As the molecular chromo-
phores were aligned in a closed ring, drastic difference in their frontier orbitals
(the well-delocalized property of highest occupied molecular orbitals [HOMO] and
localized property of lowest unoccupied molecular orbitals [LUMO]) was observed,
demonstrating the promotion in transition dipoles, mgi2, from the ground state (g) to
the intermediate state (i) (Figure 6B). It is worth noting that the serrated packing be-
tween layers exhibited much better transition dipole in comparison with those of the
eclipsed packing and the staggered packing (Figure 6E). This can be attributed to
the reduction in the undesirable non-radiative energy decay during the 2PA process,
as a result of weakening the p interactions between chromophores in adjacent layers
(Figure 3B). The serrated arrangement clearly broke the packing at the propeller-
shaped triarylamine units (electron-donor motifs) along the c axis, and also shifted
the position of benzothiadiazole functional groups (electron-acceptor motifs with
large p conjugation) to avoid overlay.

Conclusion
We show that linking molecular chromophores into 2D COFs is an effective way to
construct organic materials for two-photon absorption. The key to achieving high
2PA performance is the collective alignment of chromophores along the same direc-
tion, a unique feature offered by COFs. The growth of 2D COFs in single-crystal form
not only allowed us to visualize the arrangement of molecules within the layer but
also provided unambiguous resolution of the packing mode between layers, which
is also critical for the accumulation of their overall molecular dipoles and, therefore,
the macroscopic optical properties. In addition, the control and determination of
these packing modes provide an extra dimension for the design of COFs. COF-
606 can be made in various sizes: the intracellular distribution and uptake of COF-
606-100 nm was confirmed by confocal laser scanning microscopy in the CT26 cell
line (Figure S132). This, combined with the good biocompatibility of COF-606-
100 nm, in which the cell viability remained 90% even with the concentration of
250 mg/mL (Figure S131), revealed the possibility to use COF-606 samples in biolog-
ical applications.

Matter 2, 1049–1063, April 1, 2020 1059


EXPERIMENTAL PROCEDURES
Synthesis Procedure
The synthesis of organic building units (L-H, L-Ph, L-2MeO, L-2F, L-4F, L-BT) were
developed based on previously reported conditions with modifications to increase
the yield (Figure S3).

General Method for Synthesis and Activation of COF-601, COF-602, COF-603,


COF-604, and COF-605
A Pyrex tube with the size of 10 3 8 mm (outer 3 inner diameter) was filled with ben-
zene-1,4-diamine (25.9 mg, 0.24 mmol, 2 equiv), triarylamine derivatives (1 equiv),
o-dichlorobenzene (1.5 mL), n-butanol (1.5 mL), and aqueous acetic acid (300 mL,
6 M) sequentially. The mixture was dissolved by sonication. The solution was de-
gassed through three freeze-pump-thaw cycles. After gradual warming up to
room temperature, the tube was placed in an oven and heated at 90 C for 7 days.
The solid was then isolated by filtration and washed with tetrahydrofuran (THF),
and sequentially fluxed with THF and acetone in a Soxhlet extractor for 24 h each sol-
vent. After drying under dynamic vacuum at room temperature for 1 h, the COF sam-
ples were heated up to 80 C and kept at the same temperature for 8 h. The samples
were then cooled down to 50 C and kept for 4 h to achieve complete activation.

COF-606-100 nm
The synthesis and activation of COF-606-100 nm was similar to that of COF-605,
except that after natural warming up to room temperature, the tube was placed in
an oven and heated at 65 C for 12 h. The solid was then isolated by filtration, washed
with THF, and sequentially fluxed with THF and acetone in a Soxhlet extractor for
24 h each solvent. After drying under dynamic vacuum at room temperature for
1 h, the COF sample’s temperature was raised to 80 C and kept for 8 h, then cooled
down to 50 C and kept for 4 h to completely activate the COF sample.

COF-606-500 nm
The synthesis and activation of COF-606-500 nm was similar to that of COF-606-
100 nm, except that after natural warming up to room temperature, the tube was
placed in an oven and heated at 90 C for 7 days. The solid was then isolated by filtra-
tion, washed with THF, and sequentially fluxed with THF and acetone in a Soxhlet
extractor for 24 h each solvent. After drying under dynamic vacuum at room temper-
ature for 1 h, the COF sample’s temperature was raised to 80 C and kept for 8 h, then
cooled down to 50 C and kept for 4 h to completely activate the COF sample.

COF-606-1 mm
The synthesis and activation of COF-606-100 1 mm was similar to that of COF-606-
500 nm, except that after natural warming up to room temperature, the tube was
placed in an oven and heated by programmed temperature at a constant rate of
0.1 C–90 C and kept for 7 days. The tube was then cooled down to room tempera-
ture by programmed temperature at a constant rate of 0.1 C. The solid was isolated
by filtration, washed with THF, and sequentially fluxed with THF and acetone in a
Soxhlet extractor for 24 h each solvent. After drying under dynamic vacuum at
room temperature for 1 h, the COF sample’s temperature was raised to 80 C and
kept for 8 h, then cooled down to 50 C and kept for 4 h to completely activate the
COF sample.

BT-Amorphous
A mixture of benzene-1,4-diamine (25.9 mg, 0.24 mmol, 2 equiv) and triarylamine
derivatives L-BT (compound 4f, 88.2 mg, 0.12 mmol, 1 equiv) was dissolved in a

1060 Matter 2, 1049–1063, April 1, 2020


mixed solvent of methanol/THF (10 mL/10 mL) and aqueous acetic acid (300 mL,
6 M). The resulting mixture was heated at 65 C for 1 h under argon atmosphere.
The solid was then isolated by filtration, washed with THF, and sequentially fluxed
with THF and acetone in a Soxhlet extractor for 24 h each solvent. The material
was dried under dynamic vacuum at room temperature for 1 h. The temperature
was raised to 80 C and kept for 8 h, then cooled down to 50 C and kept for 4 h to
completely activate the COF sample.

Material Characterizations
PXRD data were collected using both synchrotron and lab-based instruments. The
X-ray diffraction data were obtained at beamline BL14B1 of the Shanghai Synchro-
tron Radiation Facility (SSRF) with a wavelength of 1.2398 Å. For data collected at
Rigaku Smartlab, a 9-kW lab-based diffractometer was operated at 45 kV and
200 mA using Cu Ka (l = 1.5418 Å) with a scan speed of 1 /min and a step size of
0.01 in 2q at ambient temperature and pressure. Simulated PXRD patterns were
calculated using the software Mercury 3.0 from the refined crystal structure based
on synchrotron data.

1
H/13C cross-polarization MAS NMR experiments were carried out at 9.4 T on a
Bruker 400 Avance III spectrometer with resonance frequencies of 399.33 and
100.42 MHz for 1H and 13C, respectively. The spectra were recorded on a 4-mm
probe with a contact time of 5 ms and a recycle delay of 2 s. The sample spinning
rate was 10 kHz. The 13C MAS chemical shifts were referenced to adamantane
(38.5 ppm).

Measurement of 2PACS Value


To accurately assess the two-photon absorption efficiency of these COFs, we used
the 2PACS value per chromophore to measure the two-photon brightness of all
the samples, which is a product of two-photon absorption section value (d) and fluo-
rescence quantum yield (h). The two-photon action cross-section (d$h) was deter-
mined by using the femtosecond fluorescence measurement technique (Section
12 of Supplemental Information), whereby each sample was measured three times
in parallel. The two-photon induced fluorescence intensity was measured in the
range from 710 to 850 nm and calibrated by rhodamine B (1 3 105 M) in methanol
as the reference. The two-photon fluorescence property of rhodamine B has been
well characterized in the literature.34 The intensities of the two-photon-induced fluo-
rescence of the reference and sample were detected at the same excitation
wavelength.

SUPPLEMENTAL INFORMATION
Supplemental Information can be found online at https://doi.org/10.1016/j.matt.
2020.01.019.

ACKNOWLEDGMENTS
H.D. acknowledges support from the National Natural Science Foundation of
China (21471118, 91545205, 91622103 and 21971199), National Key Research
and Development Program of China (2018YFA0704000), National Key Basic
Research Program of China (2014CB239203), and Innovation Team of Wuhan Uni-
versity (2042017kf0232). W.W. acknowledges support from the National Key
Research and Development Program of China (no. 2017YFA0402800) and the Na-
tional Science Foundation of China (NSFC grant U1732121). We acknowledge sup-
port from the National Key R&D Program of China (2018YFB1107700) and beamlines

Matter 2, 1049–1063, April 1, 2020 1061


BL14B1 of SSRF (Shanghai Synchrotron Radiation Facility) for providing the beam
time. We acknowledge C-EM, SPST of ShanghaiTech University (#EM02161943)
for support. We thank O. Terasaki (ShanghaiTech University), J. Caram (University
of California at Los Angeles), K.P. Loh (National University of Singapore), P. Oleyni-
kov (ShanghaiTech University), X.C. Hong, X. Zhang, J. Zeng, and F. Wang (Wuhan
University) for their invaluable discussion and help.

AUTHOR CONTRIBUTIONS
H.D. conceived the idea and led the project. L.Z. synthesized the organic linkers and
COFs. Y.Z. took the high-resolution TEM images of COFs and C.W. took the SEM
images of all the samples. L.Z., Y.Z., Y.M., and Q.L. carried out the structural charac-
terization of all the COFs. A.C., S.L., and A.M. took the STEM images of COFs. L.Z.,
W.H., and Z.L. operated the two-photon experiments. J.L. operated the power-
dependent photo-luminescence plots experiments. M.J. operated the DFT calcula-
tions. W.W. recorded the PXRD patterns of all the COFs. J.X. performed the solid-
state NMR experiments for all of the monomer and COF samples. L.Z., Y.H., Y.Z.,
X.X., J.C., and H.D. prepared the manuscript, and all authors contributed to the final
version.

DECLARATION OF INTERESTS
The authors declare no competing financial interests.

Received: April 27, 2019


Revised: January 19, 2020
Accepted: January 24, 2020
Published: February 26, 2020

REFERENCES
1. Yokoyama, D. (2011). Molecular orientation in additivity of two-photon-absorption cross 14. El-Kaderi, H.M., Hunt, J.R., Mendoza-Cortés,
small-molecule organic light-emitting diodes. sections in linear and branched squaraine J.L., Côté, A.P., Taylor, R.E., O’keeffe, M., and
J. Mater. Chem. 21, 19187–19202. super chromophores. Phys. Chem. Chem. Yaghi, O.M. (2007). Designed synthesis of 3D
Phys. 18, 16404–16413. covalent organic frameworks. Science 316,
2. Moon, C.K., Kim, K.H., and Kim, J.J. (2017). 268–272.
Unraveling the orientation of phosphors 8. Medishetty, R., Nemec, L., Nalla, V., Henke, S.,
doped in organic semiconducting layers. Nat. Samoc, M., Reuter, K., and Fischer, R.A. (2017). 15. Ding, S.Y., Gao, J., Wang, Q., Zhang, Y., Song,
Commun. 8, 791–800. Multi-photon absorption in metal-organic W.G., Su, C.Y., and Wang, W. (2011).
frameworks. Angew. Chem. Int. Ed. 56, 14743– Construction of covalent organic framework for
3. Jurow, M.J., Mayr, C., Schmidt, T.D., Lampe, T., 14748. catalysis: Pd/COF-LZU1 in Suzuki–Miyaura
Djurovich, P.I., Brütting, W., and Thompson, coupling reaction. J. Am. Chem. Soc. 133,
M.E. (2016). Understanding and predicting the 9. Drobizhev, M., Stepanenko, Y., Dzenis, Y., 19816–19822.
orientation of heteroleptic phosphors in Karotki, A., Rebane, A., Taylor, P.N., and
organic light-emitting materials. Nat. Mater. Anderson, H.L. (2005). Extremely strong near-IR 16. Diereks, C.S., and Yaghi, O.M. (2017). The
15, 85–93. two-photon absorption in conjugated atom, the molecule, and the covalent organic
porphyrin dimers: quantitative description with frameworks. Science 355, eaal1585.
4. Kim, H.M., and Bong, R.C. (2015). Small- three-essential-states model. J. Phys. Chem. B
molecule two-photon probes for bioimaging 109, 7223–7236. 17. Jin, Y., Hu, Y., and Zhang, W. (2017).
applications. Chem. Rev. 115, 5014–5055. Tessellated multiporous two-dimensional
10. Pawlicki, M., Collins, H.A., Denning, R., and covalent organic frameworks. Nat. Rev. Chem.
5. Shen, Y., Shuhendler, A.J., Ye, D., Xu, J.J., and Anderson, H.L. (2009). Two-photon absorption 1, 56–76.
Chen, H.Y. (2016). Two-photon excitation and the design of two-photon dyes. Angew.
nanoparticles for photodynamic therapy. Chem. Int. Ed. 48, 3244–3266. 18. Lohse, S.M., and Bein, T. (2018). Covalent
Chem. Soc. Rev. 45, 6725–6741. organic frameworks: structures, synthesis, and
11. Kim, H.M., and Cho, B.R. (2009). Two-photon applications. Adv. Funct. Mater. 28, 1705553–
6. Williams-Harry, M., Bhaskar, A., Ramakrishna, materials with large two-photon cross sections. 1705623.
G., Goodson, T., Imamura, M., Mawatari, A., Structure-property relationship. Chem.
Nakao, K., Enozawa, H., Nishinaga, T., and Commun. (Camb.) 153, 153–164. 19. Dogru, M., and Bein, T. (2014). On the road
Iyoda, M. (2008). Giant thienylene-acetylene- towards electroactive covalent organic
ethylene macrocycles with large two-photon 12. Feng, X., Ding, X., and Jiang, D. (2012). frameworks. Chem. Commun. (Camb.) 50,
absorption cross section and semishape- Covalent organic frameworks. Chem. Soc. Rev. 5531–5546.
persistence. J. Am. Chem. Soc. 130, 3252– 41, 6010–6022.
3253. 20. Liu, H., Chu, J., Yin, Z., Cai, X., Zhuang, L., and
13. Cote, A.P., Benin, A.I., Ockwig, N.W., O’keeffe, Deng, H. (2018). Covalent organic frameworks
7. Ceymann, H., Rosspeintner, A., Schreck, M.H., M., Matzger, A.J., and Yaghi, O.M. (2005). linked by amine bonding for concerted
Mytzel, C., Stoy, A., Vauthey, E., and Lambert, Porous, crystalline, covalent organic electrochemical reduction of CO2. Chem 4,
C. (2016). Cooperative enhancement versus frameworks. Science 310, 1166–1170. 1696–1709.

1062 Matter 2, 1049–1063, April 1, 2020


21. Li, X., Gao, Q., Wang, J., Chen, Y., Chen, Z.H., 26. Evans, A.M., Parent, L.R., Flanders, N.C., Crystallography, Vol. A, Th. Hahn, ed (Wiley),
Xu, H.S., Tang, W., Leng, K., Ning, G., Wu, J., Bisbey, R.P., Vitaku, E., Kirschner, M.S., pp. 44–54, Chapter 3.1.
et al. (2018). Tuneable near white-emissive two- Schaller, R.D., Chen, L.X., Gianneschi, N.C.,
dimensional covalent organic frameworks. Nat. and Dichtel, W.R. (2018). Seeded 30. Lazic, I., Bosch, E.G., and Lazar, S. (2016). Phase
Commun. 9, 2335–2344. growth of single-crystal two-dimensional contrast STEM for thin samples: integrated
covalent organic frameworks. Science 361, differential phase contrast. Ultramicroscopy
22. Ding, S., and Wang, W. (2013). Covalent 52–57. 160, 265–280.
organic frameworks (COFs): from design to
applications. Chem. Soc. Rev. 42, 548–568. 27. Ascherl, L., Sick, T., Margral, J.T., Lapidus, S.H., 31. Göppert-Mayer, M. (1931). Über elementarakte
Calik, M., Hettstedt, C., Karaghiosoff, K., mit zwei quantensprüngen. Ann. Phys. 9,
23. Lukose, B., Kuc, A., Frenzel, J., and Heine, T. Doblinger, M., Clark, T., Chapman, K.W., et al. 273–294.
(2010). On the reticular construction concept of (2016). Molecular docking sites designed for
covalent organic frameworks. Beilstein J. the generation of highly crystalline covalent 32. Kaiser, W., and Garrett, C.G.B. (1961). Two-
Nanotechnol. 1, 60–70. organic frameworks. Nat. Chem. 8, 310–316. photon excitation in CaF2: Eu2+. Phys. Rev. Lett.
7, 229–231.
24. Zhang, Y.B., Su, J., Furukawa, H., Yun, Y., 28. Haase, F., Troschke, E., Savasci, G., Banerjee,
33. Medishetty, R., Zareba, J.K., Mayer, D., Samoc,
Gandara, F., Duong, A., Zou, X.D., and Yaghi, T., Duppel, V., Dörfler, S., Martin, M., Grundei,
M., and Fischer, R.A. (2017). Nonlinear optical
O.M. (2013). Single-crystal structure of a J., Asbjörn, M.B., Christian, O., et al. (2018).
properties, upconversion and lasing in metal-
covalent organic framework. J. Am. Chem. Soc. Topochemical conversion of an imine-into a
organic frameworks. Chem. Soc. Rev. 46, 4976–
135, 16336–16339. thiazole-linked covalent organic framework
5004.
enabling real structure analysis. Nat. Commun.
25. Ma, T.Q., Kapustin, E.A., Yin, S.X., Liang, L., 9, 2600–2610.
34. Makarov, N.S., Drobizhev, M., and Rebane, A.
Zhou, Z.Y., Niu, J., Li, L.H., Wang, Y.Y., Su, J., Li, (2008). Two-photon absorption
J., et al. (2018). Single-crystal x-ray diffraction 29. Looijenga-Vos, A., and Buerger, M.J. (2006). standards in the 550–1600 nm
structures of covalent organic frameworks. Space group determination and diffraction excitation wavelength range. Opt. Express 16,
Science 361, 48–52. symbols. In International Tables for 4029–4047.

Matter 2, 1049–1063, April 1, 2020 1063

You might also like