-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathpaper.tex
1524 lines (908 loc) · 82.9 KB
/
paper.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
% This document is based on the RSC LaTeX template, but extensively modified,
% https://www.rsc.org/journals-books-databases/journal-authors-reviewers/author-tools-services/
% most importantly we are using LuaLaTeX instead of pdflatex in order to compile
% our large TiKZ/PGF plots (pdflatex cannot handle large memory requirements)
\documentclass[9pt,twoside,twocolumn]{article}\usepackage{knitr}
% ----------------------------------------------------------------------
% Page layout
% ----------------------------------------------------------------------
\usepackage{extsizes}
\usepackage{balance}
\usepackage[left=1.5cm,right=1.5cm,top=1.785cm,bottom=2.0cm]{geometry}
% ----------------------------------------------------------------------
% ----------------------------------------------------------------------
% Bibliography
% ----------------------------------------------------------------------
\usepackage[backend=biber,doi=true,style=chem-rsc]{biblatex}
\addbibresource{references/library.bib}
% modification: make DOI string a hyperlink in the list of references
\newcommand*\doiformat[1]{\small#1}
\DeclareFieldFormat{doi}{%
\mkbibacro{\small{DOI}}\addcolon\space
\ifhyperref{\href{https://dx.doi.org/#1}{\doiformat{#1}}}{\doiformat{#1}}%
}
% ----------------------------------------------------------------------
% ----------------------------------------------------------------------
% Fonts and symbols
% ----------------------------------------------------------------------
% \usepackage{luatextra} % also loads fixltx2e, fontspec, xunicode
\usepackage[utf8]{luainputenc}
% DO NOT use package xunicode with luatex: gives errors with tikzDevice/knitr
% https://tex.stackexchange.com/questions/401481/charter-font-shape-undefined-with-lualatex/401658
\usepackage{fontspec}
\setmainfont{XCharter}
\setromanfont{XCharter}
\setsansfont{Liberation Sans}
\setmonofont{Liberation Mono}
\usepackage{textcomp} % \textdegree, \textmu, \texttrademark and some other symbols
% ----------------------------------------------------------------------
% ----------------------------------------------------------------------
% Support packages - general
% ----------------------------------------------------------------------
\usepackage[english]{babel}
% % modification: gitinfo2 used to help keep track of manuscript versions
% \usepackage{gitinfo2}
\usepackage[compact]{titlesec}
\usepackage{sectsty}
\usepackage{fancyhdr}
\usepackage{csquotes}
\usepackage{setspace} % setstretch{} used later in the preamble
% ----------------------------------------------------------------------
% ----------------------------------------------------------------------
% Support packages - graphics, captions, footnotes, etc.
% ----------------------------------------------------------------------
\usepackage{graphicx}
\usepackage[%
format=plain,
justification=justified,
singlelinecheck=false,
font={stretch=1.125,small,sf},
labelfont=bf,
labelsep=space]{caption}
\usepackage{subcaption}
\usepackage{tikz}
% RSC template calls for float pkg, but I cannot detect that it's used
%\usepackage{float}
\usepackage{fnpos} % controls the positions of footnotes
\definecolor{cream}{RGB}{222,217,201} % RSC colour: cream
% ----------------------------------------------------------------------
% ----------------------------------------------------------------------
% Support packages - tables, lists
% ----------------------------------------------------------------------
\usepackage{booktabs}
\usepackage[inline]{enumitem}
% Reset table and figure numbering in SI (S1, S2 and so on)
% https://www.stat.berkeley.edu/~paciorek/computingTips/Customizing_numbering_pages.html
% http://bytesizebio.net/2013/03/11/adding-supplementary-tables-and-figures-in-latex/
\newcommand{\startsupplement}{%
\setcounter{table}{0}%
\renewcommand{\thetable}{S\arabic{table}}%
\setcounter{figure}{0}%
\renewcommand{\thefigure}{S\arabic{figure}}%
}
% ----------------------------------------------------------------------
% ----------------------------------------------------------------------
% Support packages - chemistry, science
% ----------------------------------------------------------------------
\usepackage{amssymb}
\usepackage{chemformula}
\setchemformula{charge-style=math}
\usepackage{siunitx}
\sisetup{%
uncertainty-mode=compact,%
reset-text-family=false,%
text-family-to-math=true,%
range-phrase=\ensuremath{\text{ to }}%
}
%https://tex.stackexchange.com/questions/56051/problem-with-options-in-declaresiunit
\DeclareSIUnit[per-mode=symbol]{\molar}{\mole\per\cubic\deci\metre}
\DeclareSIUnit[per-mode=symbol]{\micromolar}{\micro\mole\per\cubic\deci\metre}
\DeclareSIUnit{\Molar}{\textsc{m}}
\DeclareSIUnit{\vsNHE}{vs~NHE}
\DeclareSIUnit{\voltNHE}{\volt\vsNHE}
\usepackage{chemfig}
% setatomsep has been deprecated, use \setchemfig instead
\setchemfig{atom sep=2em}
% define invisible bond (for use between the charged species)
\definesubmol\nobond{-[,1.2,,,draw=none]}
% ----------------------------------------------------------------------
% ----------------------------------------------------------------------
% Packages that need to be called last: hyperref, cleveref etc.
% ----------------------------------------------------------------------
% do not use [pagebackref]{hyperref}, not stable, gives errors
\usepackage{hyperref}
% modication: cleveref is indispensable
\usepackage[nameinlink]{cleveref}
% ----------------------------------------------------------------------
% ----------------------------------------------------------------------
% R SETUP
% ----------------------------------------------------------------------
% ----------------------------------------------------------------------
% ----------------------------------------------------------------------
% R LOAD PACKAGES
% ----------------------------------------------------------------------
% ----------------------------------------------------------------------
% !Rnw root = paper.Rnw
% !Rnw root = paper.Rnw
\IfFileExists{upquote.sty}{\usepackage{upquote}}{}
\begin{document}
\pagestyle{fancy}
\thispagestyle{plain}
\fancypagestyle{plain}{
% Simple header without header graphics
\renewcommand{\headrulewidth}{0pt}
}
\makeFNbottom
\makeatletter
\renewcommand\LARGE{\@setfontsize\LARGE{15pt}{17}}
\renewcommand\Large{\@setfontsize\Large{12pt}{14}}
\renewcommand\large{\@setfontsize\large{10pt}{12}}
\renewcommand\footnotesize{\@setfontsize\footnotesize{7pt}{10}}
\makeatother
\renewcommand{\thefootnote}{\fnsymbol{footnote}}
\renewcommand\footnoterule{\vspace*{1pt}%
\color{cream}\hrule width 3.5in height 0.4pt \color{black}\vspace*{5pt}}
\setcounter{secnumdepth}{5}
\makeatletter
\renewcommand\@makefntext[1]%
{\noindent\makebox[0pt][r]{\@thefnmark\,}#1}
\makeatother
\renewcommand{\figurename}{\small{Fig.}~}
\sectionfont{\sffamily\Large}
\subsectionfont{\normalsize}
\subsubsectionfont{\bf}
\setstretch{1.125}
\setlength{\skip\footins}{0.8cm}
\setlength{\footnotesep}{0.25cm}
\setlength{\jot}{10pt}
\titlespacing*{\section}{0pt}{4pt}{4pt}
\titlespacing*{\subsection}{0pt}{15pt}{1pt}
\fancyfoot{}
% \fancyfoot[CO]{\vspace{-7.1pt}\hspace{13.2cm}}
% \fancyfoot[CE]{\vspace{-7.2pt}}
\fancyfoot[RO]{\footnotesize{\sffamily{\textbar \hspace{2pt}\thepage}}}
\fancyfoot[LE]{\footnotesize{\sffamily{\thepage~\textbar}}}
\fancyhead{}
\renewcommand{\headrulewidth}{0pt}
\renewcommand{\footrulewidth}{0pt}
\setlength{\arrayrulewidth}{1pt}
\setlength{\columnsep}{6.5mm}
\makeatletter
\newlength{\figrulesep}
\setlength{\figrulesep}{0.5\textfloatsep}
\newcommand{\topfigrule}{\vspace*{-1pt}%
\noindent{\color{cream}\rule[-\figrulesep]{\columnwidth}{1.5pt}} }
\newcommand{\botfigrule}{\vspace*{-2pt}%
\noindent{\color{cream}\rule[\figrulesep]{\columnwidth}{1.5pt}} }
\newcommand{\dblfigrule}{\vspace*{-1pt}%
\noindent{\color{cream}\rule[-\figrulesep]{\textwidth}{1.5pt}} }
\makeatother
\twocolumn[
\begin{@twocolumnfalse}
\vspace{3cm}
\sffamily
\begin{tabular}{m{4.5cm} p{13.5cm} }
\textsf{\footnotesize DOI: 10.1021/acs.jpcc.9b11229} &
\noindent\LARGE{\textbf{Optical properties of ZnO quantum dots and effect of size on their photocatalytic activity$^\dag$}}\\
\vspace{0.3cm} & \vspace{0.3cm}\\
{\footnotesize\sffamily\textit{J.\ Phys.\ Chem.\ C} 2020, 124, 11} &
\noindent\large{Taha Ahmed\textit{$^{a}$} and Tomas Edvinsson\textit{$^{b\ddag}$}} \\
\hspace{1.09in} & \noindent\normalsize{
Zinc oxide is a well-known metal oxide and with its direct optical band gap, it
provides a promising alternative to titanium oxide in photocatalytic applications.
\ch{ZnO} is here studied as quantum dots (QDs) where ultra-small nanoparticles of
\ch{ZnO} show optical quantum confinement with a band gap opening for particles below
\qty{9}{\nm} in diameter from the shift of the band edge energies.
The optical properties of growing \ch{ZnO} quantum dots are determined with
Tauc-analysis and a system of QDs for treatment and degradation of distributed
threats is analysed using an organic probe molecule, methylene blue.
The effect of optical properties of the QDs and the kinetics of dye degradation
are quantified for low-dimensional ZnO materials in the range of \qtyrange{3}{7}{\nm}
and show a substantial increase in photocatalytic activity compared to larger
\ch{ZnO} particles.
This is attributed to a combined effect from the increased surface area as well
as a quantum confinement effect that goes beyond the increased surface area.
The results show a significantly higher photocatalytic activity for the QDs between
\qtyrange[range-phrase=--]{3}{6}{\nm} with a complete decolouration of the model
compound, while QDs from \qty{6}{\nm} and upwards in diameter show signs of competing
reduction reactions.
Our study demonstrates that ultra-small \ch{ZnO} particles have a reactivity beyond
that which is expected due to their increased surface area and also show
size-dependent reaction pathways which introduces the possibility for a size
selective catalysis.%
}\\
\end{tabular}
\end{@twocolumnfalse} \vspace{0.6cm}
]
\rmfamily
\section*{}
\vspace{-1cm}
\footnotetext{\textit{$^{a}$~Department of Chemistry--{\AA}ngstr{\"o}m, Structural chemistry, Uppsala university.}}
\footnotetext{\textit{$^{b}$~Department of Engineering Sciences, Solid state physics, Uppsala university, Sweden; tomas.edvinsson@angstrom.uu.se}}
\footnotetext{%
\dag~The published paper and its supplementary information is available online:\newline
\url{https://pubs.acs.org/doi/full/10.1021/acs.jpcc.9b11229}%
}
% ----------------------------------------------------------------------
% Custom LaTeX commands
% ----------------------------------------------------------------------
% optical absorbance
\newcommand{\absorbance}{\ensuremath{A}}
% ----------------------------------------------------------------------
% references inside main paper (separated from refs in SI)
\begin{refsection}
% Code (data analysis)
% !Rnw root = paper.Rnw
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%% small NP
%%% NP initial size: SMALL
%%% NO STIRRING
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%% small NP
%%% NP initial size: SMALL
%%% NO STIRRING
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%% medium NP
%%% NP initial size: MEDIUM
%%% NO STIRRING
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%% large NP
%%% NP initial size: LARGE
%%% NO STIRRING
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%% large NP
%%% NP initial size: LARGE
%%% NO STIRRING
% Body text
% !Rnw root = paper.Rnw
\section{Introduction}
Zinc oxide is an abundant and familiar material that has applications spanning from UV-absorbing additive in rubber \cite{Heideman2005} and sunscreens \cite{King2005,Serpone2007}, to high-tech applications in chemical sensing \cite{Rambu2013,Cai2013}, spintronics \cite{Sharma2004}, and photovoltaics \cite{Quintana2007,Boschloo2006}.
The wide interest in low-dimensional \ch{ZnO} for optoelectronic applications is fuelled by its high carrier mobility \cite{Yoshikawa1997,Look1998,Kaidashev2003}, its direct wide band gap ($E_\text{g}$ $\sim\qty{3.3}{\eV}$), and a large exciton binding energy of \qty{60}{\meV} \cite{Ozgur2005} allowing room-temperature exciton emission.
There are also intriguing size effects of \ch{ZnO} on the nano-scale below \qty{9}{\nm} in diameter, such as vibrational quantum confinement \cite{Raymand2012}, discontinuous exciton emission decay rate \cite{Jacobsson2014}, and quantum confined Stark effects \cite{Jacobsson2014a} that show that more is to be discovered even in this seemingly well-known material.
Here, we focus on the optical and photochemical properties of \ch{ZnO} quantum dots with a special focus towards their possible application as photocatalysts in colloidal suspension for the degradation of distributed environmental threats.
The formal quantum efficiency of photocatalysis is higher in smaller \ch{ZnO} quantum dots compared to larger ones and show an interplay between the accessible surface area and their ability to absorb photons in the solar spectrum \cite{Jacobsson2012} and is likely also related to the interaction and thus the weakening of excitons with charged surface states \cite{Jacobsson2014}.
Here it is interesting to compare the properties of wurtzite \ch{ZnO} with that of the more extensively studied anatase \ch{TiO2}.
Zinc oxide has a markedly higher absorption coefficient due to its direct band gap, almost the same conduction and valence band edges (\qty{-0.5}{\volt} and \qty{+2.8}{\voltNHE} at \mbox{pH\,5}) as \ch{TiO2} and the advantage of a higher charge carrier mobility \cite{Boschloo2006} showing more or less the same driving force for the reactions but a more effective absorption of light and improved carrier transport.
There are also several studies reporting a more effective photocatalytic degradation of organic dyes for ZnO compared to \ch{TiO2} \cite{Sakthivel2003,Muruganandham2006,Yan2009a}.
Although \ch{ZnO} does not have any intrinsic toxicology, the nanoscale size of the particles raises the issue of nano-toxicity where particles smaller than \qty{100}{\nm} typically introduce surface reactivities causing nanoscale toxicity \cite{Maynard2011, Maynard2011a}.
However, the higher reactivity for nanoscale materials is a desirable property for the degradation of environmental threats where the high surface areas and the ability to penetrate into porous materials, such as soot particles and oils, are advantageous.
If a dynamical nanoscale system, such as the slowly growing nanoparticles in this study, is applied towards photocatalysis the higher reactivity can be used when the particles are small and the system will then naturally be deactivated as the particles grow larger or aggregate among themselves or with their environment during the later stages of the process.
The system with the photoactive \ch{ZnO} shows slow system growth and coalescence into aggregates and natural sedimentation to the bottom of the contaminated water container or, if applied at larger scale, to the bottom of the pond or sea to become a part of the natural sediment, thus deactivating its nano-toxicity.
A key focus of our approach is to investigate the properties of low-dimensional \ch{ZnO} under growth by studying their optical and photocatalytic activity in the quantum confined regime ($d<\qty{9}{\nm}$) where the band gap of the particles changes during their growth with eventual aggregation and sedimentation for larger particles.
The United Nations World water development report \cite{WWAP2017} identifies the effective removal of emerging pollutants
(such as pharmaceuticals, fragrances, pesticides and herbicides)
as one of the main challenges for the European region of the world, and as a large concern in other regions, where environmental regulations and laws may still lag behind industrial practices.
When it comes to organic compounds, one effective approach to degrade such pollutants is to utilise the fact that the energy in UV-photons is high enough to break covalent bonds in molecules.
For some unstable molecules such direct illumination is enough to degrade them, but generally a photocatalyst is needed to overcome the activation barriers and significantly improve the degradation kinetics.
The effect is mediated \textit{via} various Reactive Oxygen Species (ROS), such as singlet oxygen \ch{^1O2}, superoxide radicals \ch{O2^{.-}}, hydroperoxyl radicals \ch{HO2^.}, hydroxyl radicals \ch{^.OH}, and hydrogen peroxides \ch{H2O2} that are generated at the metal oxide surface upon UVA (\qtyrange{315}{400}{\nm}) irradiation in aqueous solutions \cite{Tachikawa2007}.
For \ch{ZnO}, the photo-generated holes in the valence band are generated with a higher rate compared to the indirect band gap semiconductor \ch{TiO2}, while having the photogenerated hole still sufficiently low in energy and thus highly oxidising (see~\cref{fig:schematic-and-solarlamp}).
When water is present at the semiconductor surface, the photo-generated hole generates hydroxyl radicals (\ch{^.OH}) \cite{Herrmann2005,Chen2007d} that act as a strong non-selective oxidant.
The end result is often a complete mineralisation of the undesired chemicals into \ch{CO2}, water and mineral acids.
In this paper, we show that a small change in \ch{ZnO} nanoparticle diameter leads to a large change in photocatalytic activity. The observed behaviour is a result of the size-dependent change of energy levels and the subsequent optical effect in the quantum-confined semiconductor system in aqueous suspension.
\section{Methods}
A series of methylene blue aqueous solutions of with concentrations of \qtylist{0.5;1;2;5;10;20;50;100}{\micromolar} were volumetrically prepared from methylene blue (powder, VWR Prolabo, CAS 61-73-4) weighed on a Mettler AT261 balance with \qty{0.01}{\mg} resolution, quantitatively transferred and diluted to the mark in volumetric glassware with deionised water (for volumes and dilution factors, see \cref{tabsi:mb-stock-solutions}).
The dye solutions were stored in bluecap glass bottles wrapped in aluminium-foil and stored in the dark.
\ch{ZnO} nanoparticle suspensions were synthesised by mixing zinc acetate dihydrate, \ch{Zn(OAc)2 * 2 H2O}, Fluka, $\geqslant\qty{99.5}{\percent}$, and lithium hydroxide monohydrate, \ch{LiOH * H2O}, Sigma-Aldrich, $\geqslant\qty{99.0}{\percent}$, both dissolved in ethanol, \qty{99.5}{\percent}, following previously published routes \cite{Jacobsson2011,Meulenkamp1998,Spanhel1991}.
In brief, each precursor was dissolved in \ch{EtOH} and diluted to \qty{50}{\ml} (\textit{cf.} \cref{tabsi:precursor-solutions}).
The precursor solutions were then quantitatively mixed and the time was noted (this represented the start of nanoparticle nucleation and growth).
Our modifications to previous work included heating the \ch{LiOH} solution to around \qty{45}{\celsius} while flushing the head space with \ch{N2} to both increase the solubility of lithium hydroxide and decrease the chance of formation of the carbonate.
Just prior to mixing both precursor solutions were cooled in an ice bath for a few minutes.
In all PC experiments, one part of the \ch{ZnO(EtOH)} NP suspension was mixed with nine parts of \qty{10}{\micromolar} \ch{MB(aq)} directly into a \qty{10}{\mm} PMMA cuvette (typically $\qty{0.2}{\ml}+\qty{1.8}{\ml}$, so the cuvette contained \ch{H2O}:\ch{EtOH} 90:10 by volume), which after an initial thorough but fast mixing was with a solar simulator (\qty{100}{\mW\per\square\cm}, Oriel/Newport class~A solar simulator with an AM1.5G optical filter) while recording a full UV/Vis/NIR spectrum perpendicular to the direction of illumination each minute for the duration of the experiment.
The UV/Vis/NIR absorption measurements were performed using an OceanOptics HR-2000+ spectrometer with deuterium and halogen light sources.
In all our measurements, a spectrum with \num{2048} evenly distributed points from \qtyrange{190}{1100}{\nm} were sampled with an integration time of \qty{1}{\ms} (each saved spectrum averaged \num{10} such measurements) using fibre-optic cables from the lamp to the cuvette holder and back to the spectrometer, such that the face of the cuvette being illuminated by the solar lamp was unobstructed.
We used PMMA cuvettes ($l=\qty{1}{\cm}$) with transparent walls and the top of the cuvette was left open to atmosphere during PC experiments. We observed some evaporative loss during a typical PC experiment (duration approximately \qty{150}{\minute}) but it was small and outweighed by the need to allow air into the cuvette.
Band gap analysis was performed using Tauc plots, where the square of the absorption was plotted as a function of photon energy.
For a direct band gap semiconductor like \ch{ZnO}, the band gap is then given as the intercept of the fitted line with the abscissa. More details on the Tauc fit procedure can be found in the SI.
For all recorded spectra, the methylene blue absorption band (roughly \qtyrange{500}{800}{\nm}) was baseline-fitted with a cubic spline function and peak-fitted with two Gaussian kernels in an automatic fashion using the scripting-capabilities of Fityk~1.3.1 \cite{Wojdyr2010}.
The baseline correction was needed because the baseline of the spectra increased continuously over the course of an experiment due to increased scattering, in turn due to processes involved with agglomeration of the nanoparticles with each other and with the organic species in solution over the course of an experiment.
\section{Results and discussion}
Under the scheme proposed in this paper, a suspension of \ch{ZnO} nanoparticles in their native growth solution would be added to a polluted water surface and with exposure to sunlight photocatalyse the breakdown of the organic pollutant followed by the aggregation and subsequent sedimentation of the catalyst particles themselves, effectively rendering them inert.
Here we spectrometrically track both the \ch{ZnO} band edge and the methylene blue (MB) absorption band \textit{in-situ} under simulated solar illumination to gain insights into both how the optical properties of the QDs change, the effect of growth upon dilution and their photocatalytic activity during the process.
\subsection{Optical properties of low-dimensional \ch{ZnO}}
\begin{figure}[tb]
\centering
\begin{subfigure}[b]{\linewidth}
\caption{}
\label{fig:schematic}
\includegraphics[width=\linewidth]{images/nanoparticle-redox}
\end{subfigure}\\%
\begin{subfigure}[b]{\linewidth}
\caption{}
\label{fig:solarlamp}
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=3.40in]{figure/fig-solarlamp-1}
}
\end{knitrout}
\end{subfigure}
\caption{(\subref{fig:schematic}) Schematic of the different processes occurring on a catalyst nanoparticle's surface during photocatalysis in water, along with an energy diagram of the electronic states in low-dimensional \ch{ZnO}.
(\subref{fig:solarlamp}) The ASTM AM1.5G solar spectrum (here shown only up to \qty{1000}{\nm}) overlaid with: UV absorbance of PMMA cuvette (grey area, \textit{c.f.} \cref{figsi:cuvette-PMMA-abs}), with the coloured areas showing the wavelength region of the solar spectrum absorbable by our smallest (yellow) and largest (blue) \ch{ZnO} NPs, respectively.
The areas shaded in darker yellow and darker blue underline part of the spectrum of our laboratory's AM1.5G solar simulator (which outputs less UV and more VIS than the ASTM reference).}
\label{fig:schematic-and-solarlamp}
\end{figure}
The optical properties of ultra-small \ch{ZnO} particles can theoretically be determined by solving the Schrödinger equation for the first excited state for a particle in a spherical well with radius $R$ for the kinetic energy and a higher frequency dielectric screening approach for the potential energy giving \cite{Brus1984,Jacobsson2011,Edvinsson2018}
\begin{equation}
E_\text{g} = %
E_\text{g,\,bulk} +
\frac{\hbar^{2}\pi^{2}}{2R^2} \left(\frac{1}{m_\text{e}^{*}} + \frac{1}{m_{h}^\text{*}}\right) %
- \frac{1.8e^2}{\varepsilon R} +
\frac{e^2}{R}\sum_{n=1}^{\infty} \alpha_\text{n} \left(\frac{r_\text{e} + r_\text{h}}{R}\right)^{2n}
\end{equation}
where $E_\text{g,bulk}$ is the band gap of the bulk material, $\hbar$ is Planck's constant, $\varepsilon$ is the dielectric constant of \ch{ZnO}, $e$ is the elementary charge, and $m_\text{e}^{*}$ and $m_\text{h}^{*}$ are the effective masses of the electron and hole, respectively.
The second term is the increased kinetic energy from the localisation of the electron-hole pair confined within a sphere with radius $R$ and scales as $R^{-2}$.
The third term is the Coloumb attraction in a screened environment and scales as $R^{-1}$.
The last term describes the average polarisability for an electron and hole at position $r_\text{e}$ and $r_\text{h}$, respectively, with a polarisability $\alpha_\text{n} = \frac{(\varepsilon_\text{rel} - 1)(n + 1)}{\varepsilon_2(\varepsilon_\text{n} + n + 1)}$ where $\varepsilon_\text{rel} = \frac{\varepsilon_2}{\varepsilon_1}$ is the relative dielectic constant of a material with $\varepsilon_2$ in an environment with the dielectric constant $\varepsilon_1$-.
This polarisation term goes to zero for a localised excitation in a screened environment, that is, if $r_\text{e} + r_\text{h} \ll R$. \cite{Brus1984}
This gives a physical justification for a functional dependence between the band gap, $E_\text{g}$, in \unit{\eV} and the particle diameter, $d$, in \unit{\nm} as $E_\text{g} = C_1 + \frac{C_2}{d} + \frac{C_3}{d^2}$ where the coefficients $C_n$ are to be found from experimental data.
\textcite{Jacobsson2011} found that $C_1=\qty{3.301}{\eV}$, $C_2=\qty{0.293}{\eV\nm}$, and $C_3=\qty{3.94}{\eV\nm\squared}$ for low-dimensional \ch{ZnO} using growth conditions similar to ours.
One can note that the use of bulk band gap $E_\text{g}=\qty{3.301}{\eV}$ intrinsically includes the exciton absorption in \ch{ZnO} as this cannot be deconvoluted from the experimental absorption onset at room temperature.
The mean diameter of the \ch{ZnO} nanoparticle ensemble is given by rearranging the previous formula into
\begin{equation}\label{eq:diameter}
d = \frac{2C_3}{-C_2 + \sqrt{C_2^2 - 4 C_3 (C_1 - E_\text{g})}}\text{.}
\end{equation}
To quantify the photocatalytic properties of the distributed particles, and in order to get a better idea of how the material would perform under natural sunlight, we used an AM1.5G solar simulator lamp (\textit{c.f.} \cref{fig:solarlamp}). \cite{ASTMG17303}
The large band gap of \ch{ZnO} severely restricts the number of solar photons the material can absorb, and this is further exacerbated initially by the large band gap of the smallest-diameter particles.
At $E\geqslant\qty{3.78}{\eV}$ only \qty{0.14}{\percent} of the AM1.5G photon flux is available to the \ch{ZnO} photocatalyst, and at the limit of the bulk band gap ($E\geqslant\qty{3.30}{\eV}$) \qty{1.19}{\percent} is available, or about 8~times more photons.
This can be expressed as a relative (external) quantum efficiency (QE), which is the ratio of the available photon flux at the current band gap divided by the maximum available photon flux, $\Phi$, at the bulk band gap (which corresponds to unity when $E = E_\text{g}$).
\begin{equation}\label{eq:relative-QE} % photon flux
\mathrm{QE}_\text{rel} = %
\frac{\Phi(E_\text{g})}{\Phi(E_\text{g} = E_\text{g,\,bulk})} \times d^2
\end{equation}%
We should also note that while the band gap of the \ch{ZnO} nanoparticles (NPs) puts a lower limit on the photon energies our photocatalyst can potentially absorb, the absorption onset of the cuvette itself (made of polymethylmetacrylate, PMMA) applies an upper limit where absorption starts at \qty{280}{\nm}, which coincides with the high-energy terminus of the AM1.5G spectrum. But this limit could easily be circumvented by running the photocatalysis in an open (cuvette-less) system under natural sunlight.
The surface area ($A=4\pi r^2$) affects the relative QE by a factor \num{10} (for NP diameters from approximately \qtyrange{4}{9}{\nm}).
Looking at each factor in the relative QE, we may note that as the band gap increases, the relative QE decreases, and as the surface area (\textit{i.e.}, diameter) increases, the relative QE increases.
In total, the relative QE thus increases as the nanoparticles grow, from their spectral absorption increase, but at the same time their available surface area decreases.
In terms of magnitude, and based on the ideal conditions we have implicitly assumed above (perfect spheres, smooth surfaces, etc.), both factors are well matched, as evident.
So the smallest \ch{ZnO} NPs suffer the smallest surface area \emph{and} absorb the least amount of solar photons, while the largest \ch{ZnO} NPs enjoy the largest surface area and absorb an amount of solar photons approaching the bulk-limit. Based on this, we would not expect nanoparticles to confer any advantage. But they do, and it is primarily related to the higher catalytic efficiency seen for the smaller particles, thought to be closely related to the increased surface reactivity due to an increased incidence of surface defects and a higher degree of boundaries/edges between different crystalline surfaces.
\subsection{Methylene blue}
To probe the photoreactivity of the \ch{ZnO} nanoparticles we added them to an aqueous solution of \qty{10}{\micromolar} methylene blue, which is a cationic organic dye commonly employed as a model pollutant in photocatalysis \cite{Schubert1955} and belongs to the group of thiazine dyes.
\Cref{fig:MB-abs-coefficient} shows the chemical structure of methylene blue (bis-3,9-dimethylamino phenazothionium chloride), with the positive charge assigned to the thiazine sulfur.
The cation is a heterocyclic aromatic (note that the amine nitrogens are part of the aromatic system and cannot rotate), with a single formal positive charge delocalised across the $\pi$-system.
Although the use of methylene blue as a photocatalytic probe is widespread, in our opinion its application is complicated by several factors: %
\begin{enumerate*}[label=\alph*)]%inline list
\item oligomerisation in aqueous solution leading to metachromasy and non-trivial UV/Vis band assignment,
\item solvent effects (due to mixing EtOH into the MB aqueous solution),
\item self-degradation of MB under illumination, and
\item photobleaching/decolourisation by a two-electron reduction breaking the aromaticity but otherwise leaving the molecule structurally intact (forming leuco methylene blue).
\end{enumerate*}
MB cations are also known to form di-, tri- and possibly even higher oligomers in aqueous solution (the \ch{MB+} stack with the $\pi$-systems facing each other, thus forming multi-layered \enquote{sandwiches}) likely driven by hydrophobic forces and dispersion forces between the planar $\pi$-systems. \cite{Cenens1988,Braswell1968,Schubert1955}
The absorption band of the monomer is commonly reported at \qty{665}{\nm}, \cite{Heger2005,Cenens1988} and oligomerisation is indicated by other absorption bands: dimer at around \qty{610}{\nm} \cite{Heger2005,Jockusch1995,Ruprecht1984} and trimer at around \qty{565}{\nm} \cite{Heger2005} (all reported at RT). Naturally, the degree of oligomerisation in aqueous solution depends on the total MB concentration:
\textcite{Heger2005} observed dimers above \qty{1E-4}{\molar}, while \textcite{Jockusch1995} reported dimerisation above \qty{1E-5}{\molar}. The latter value agrees well with our spectrophotometric observations.
Since a different solvent environment is expected to affect the rates and equilibrium positions of chemical reactions, as well as the position and intensity of UV/Vis absorption bands, \cite{Reichardt1994} we plotted the absorption coefficient of MB in aqueous solution and in \ch{H2O}:\ch{EtOH} (90:10) solution based on the measured absorbance spectra (\qtyrange{250}{900}{\nm}) for seven concentrations between \qtyrange[range-phrase=\ensuremath{\text{ and }}]{0.5}{50}{\micromolar} (\cref{fig:MB-abs-coefficient}) using Lambert-Beers law, $A(\lambda,c_n) = \alpha lc_n$
where $A$ is the measured absorbance, $\alpha$ is the absorption coefficient, $l$ is the optical path length, and $c_\text{n}$ the concentration for species $n$.
We used peak fitting with Gaussian profiles and found the monomer $n$-$\pi^\ast$ band at \qty{666}{\nm} (labelled \textsf{1} in \cref{fig:MB-water-fitted}), the \num{0}-\num{1} vibronic transition of the monomer band at \qty{627}{\nm} (\textsf{2}), the dimer band at \qty{608}{\nm} (\textsf{3}), and the monomer $\pi$-$\pi^\ast$ UV band at \qty{292}{\nm}.
All observed bands are listed in \cref{tabsi:MB-fitted}.
It has long been known that the dimerisation of MB disappears in solvents of lower dielectric constant (assuming constant temperature), \cite{Lewis1943} and our results show that dimerisation is completely suppressed in the \ch{H2O}:\ch{EtOH} 90:10 solvent (as evident by the suppressed dimer band, \textit{cf.} \cref{fig:MB-ethanol-fitted}).
This suggests that EtOH preferentially solvates the MB cation, which is reasonable considering the lower hydrogen-bonding ability of \ch{EtOH} compared to \ch{H2O}.
Furthermore, the MB spectrum in the presence of EtOH also serves to confirm the validity of our peak assignment in water, by clearly demonstrating that the \qty{627}{\nm} shoulder (\textsf{2}) is the \num{0}-\num{1} vibronic transition and not the dimer band, and also demonstrating that the two features overlap.
This point has been a cause for some confusion in the literature, where some authors have assigned the shoulder as either the vibronic band \emph{or} the dimer band, \cite{Heger2005,Parkanyi1993,Cenens1988,Bergmann1963} whereas in reality both features may have contributed to the observed shoulder (dependent on MB concentration, solvent composition, and temperature).
We measured the self-degradation of MB in \ch{H2O} and \ch{H2O}:\ch{EtOH} solution under 1~Sun illumination (see \cref{figsi:MB-self-degradation-spectra}) and found that MB was quite stable, as expected.
The self-degradation rate was not completely linear over time, but nearly so, and remained consistent over more than \qty{200}{\minute}. The self-degradation rate was slightly slower in the EtOH-containing solvent.
The final point that may complicate the analysis of MB degradation from UV/Vis absorption spectra is the reduction of the MB cation to its leuco (colourless) form, which can be expressed by the formula: \cite{Bard2008,Hallock2003,Resch1990}
\begin{equation}\label{eq:MB-reduction}
\ch{MB+ + H+ + 2 e- <=> MBH}
\end{equation}
We may note that the formation of LMB proceeds via the formation of a semi-reduced MB radical \ch{MB + e^-_{CB} -> MB^{.-}} that subsequently disproportionates \ch{2 MB^{.-} -> MB+ + MBH}. \cite{Mills1999}
If the electrons come directly from the CB of the semiconductor, the reduced dye is expected to adhere to the \ch{ZnO} site, effectively poisoning the catalyst. \cite{Baruah2009,Bora2016}
The electrons could also come from a subsequently created oxygen radical, a sacrifical electron donor (SED), or acetate or even the ethanol solvent itself, but this process would also compete with the SEDs scavenging VB holes. \cite{DeTacconi1997}
This photobleaching of MB can be reversed if \ch{O2} is bubbled into the solution: \ch{2 MBH + O2 -/> 2 MB+ + 2 H2O}. \cite{Mills1999}
\textcite{DeTacconi1997} claims that MB recolouration can proceed even in the absence of dissolved oxygen, principally by electron transfer from LMB into the CB (or nearby trap states) of the oxide, although this is contested by \textcite{Mills1999}.
In any case, it is clear that MB can switch from a coloured (aromatic) state to a leuco state (with broken aromaticity) in a photoreduction process, and the leuco form has no absorption bands in the visible spectral range.
\begin{figure}[tbp]
\centering
\begin{subfigure}[b]{\linewidth}
\caption{}
\label{fig:MB-abs-coefficient}
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=3.40in]{figure/fig-MB-abs-coefficient-1}
}
\end{knitrout}
\end{subfigure}\\
\begin{subfigure}[b]{0.49\linewidth}
\caption{}
\label{fig:MB-water-fitted}
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=1.75in]{figure/fig-MB-water-fitted-1}
}
\end{knitrout}
\end{subfigure}%
\hfill%
\begin{subfigure}[b]{0.49\linewidth}
\caption{}
\label{fig:MB-ethanol-fitted}
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=1.75in]{figure/fig-MB-ethanol-fitted-1}
}
\end{knitrout}
\end{subfigure}%
\caption{(\subref{fig:MB-abs-coefficient}) Absorption coefficients for methylene blue in \ch{H2O} and \ch{H2O:EtOH} (90:10 by volume) solutions, determined by experimentally measuring absorbance spectra for seven concentrations between \qtyrange[range-phrase=\ensuremath{\text{ and }}]{0.5}{50}{\micromolar}.
The chemical structure of methylene blue is shown as inset.
(\subref{fig:MB-water-fitted}, \subref{fig:MB-ethanol-fitted}) The components of the peak fit in the visible range are shown: absorption coefficients (black dots), sum of Gaussian profiles (blue line), Gaussian profiles (red line, numbered), and residual (azure line).
Parameters of all fitted bands are listed in \cref{tabsi:MB-fitted}.}
\label{fig:MB-abs-coefficients}
\end{figure}
In our estimation, and considering the aim of the present work, which is to evaluate the size-dependence of the photoreactivity of \ch{ZnO} nanoparticle suspensions during growth, MB is the most suitable dye as the ZnO QD properties also has to be compared to the vast literature of previous studies for various materials using MB as the probe.
Another important consideration is that the dye should not have absorption bands overlapping the \ch{ZnO} band edge, which is wider in our case due to its dependence on the particle diameter.
\subsection{Optical and photocatalytic properties}
As the majority solvent is water and the minority solvent is ethanol, and their miscibility is expected to be complete, we can treat the solvent system as homogeneous after the initial mixing.
But on a more immediate time scale, we have water containing only methylene blue, while the ethanol phase contains a more complex mixture: precursors, namely \ch{Li+}, \ch{OH-}, \ch{Zn^{2+}}, \ch{CH3COO-}, along with growing ZnO nanoparticles and some water (from crystalline water in \ch{LiOH} and \ch{Zn(OAc)2}).
The hydroxide and zinc ions will form the \ch{ZnO} nanoparticles, with the \ch{Li+} remaining in solution (or coordinated to the predominantly negatively charged \ch{ZnO} surfaces).
The acetate ions are assumed to remain in solution.
With the cuvette open to atmosphere oxygen is free to dissolve in the solvent, but since \ch{O2} was not actively supplied during photocatalysis the oxygen content will at most be that of the oxygen solubility at room-temperature, namely \qty{9.1}{\mg\per\litre} (equivalent to \qty{0.28}{\mmol\per\litre} or approximately 10 molecules of dissolved \ch{O2} per million molecules of \ch{H2O}) \cite{Hitchman1978,water2018}.
As we have discussed above, dye decolourisation is not always equal to decomposition: it can also occur when the dye is reduced, oxidised or adsorbed on the photocatalyst or the walls of the reactor.
A complete mineralisation of MB can require up to 70 consecutive oxidation steps where the formal quantum efficiency also can be \qty{0.1}{\percent} which means that only one out of \num{1000} photons will produce ROS that are photocatalytically active \cite{Jacobsson2012} for decolourisation.
Considering that reduction can interfere in both the initial de-colourisation as well as in subsequent steps, this makes the exact quantification of the full mineralisation complex.
In the present study, we utilise the de-colourisation kinetics as a relative measure of the photocatalytic efficiency as commonly done in the litterature but it has to be remembered that this intrinsically only quantifies the decolourisation process and not the spectrally invisible steps.
\Cref{fig:MB-N04H-small-nostir-photodegradation} shows the recorded optical absorption spectra (with lines drawn with increasingly lighter shades of blue with progressing time $t$, and with the the first and last recorded spectrum highlighted in red) of a suspension of \ch{ZnO} QDs growing from a diameter of approximately \qty{3.2}{\nm} to \qty{6}{\nm} in the presence of MB and under solar illumination.
The plot is drawn as a Tauc plot with the ordinate squared (since \ch{ZnO} has a direct band gap) and a straight line fitted to the \ch{ZnO} band edge of each spectrum, but with the $y$-axis square-root-transformed in order to not suppress the MB absorption peak visually (in comparison with the otherwise visually dominating band edge region).
The clear upwards shift of the baseline over time is due to the increased scattering losses incurred from the increasing turbidity in the cuvette, which were likely due to the aggregation and interlinking of the growing \ch{ZnO} particles with the organic species in solution (the solution in the cuvette was observed to become more viscous). During the experiment, the cuvette contained methylene blue in a \ch{H2O}:EtOH 90:10 solution (by volume) in the presence of our ZnO nanoparticles (in suspension). The cuvette was under 1~Sun illumination.
The plot displays the shifting absorbance onset (due to the changing quantum confinement) of the growing \ch{ZnO} QDs, and the concurrent bleaching of the MB band.
The top inset (\cref{fig-a:MB-N04H-small-nostir-photodegradation}) of the same figure shows the decrease in absorbance of the main MB band with time, which shows that complete decolouration occurred after approximately \qty{100}{\minute}.
The straight line (yellow) fitted to each band edge shows the result of the Tauc analysis of each spectrum, and the calculated optical band gap (\textit{i.e.}, the intersection of each yellow line with the abscissa) is plotted in the bottom inset (\cref{fig-b:MB-N04H-small-nostir-photodegradation}) along with the calculated particle diameter (both vs time).
The Tauc analysis shows that the already aged \ch{ZnO} nanoparticles in suspension grow slowly during the experiment from \qty{3.2}{\nm} to close to \qty{6}{\nm} during \qty{140}{\minute} (2h and 20 minutes). Note that the band edge gradually flattens out with time (see main plot), rendering the Tauc analysis more incertain after around \qty{20}{\minute}.
The methylene blue absorption peak was analysed by peak fitting the two main bands (compare the highlighted box in the plot) and finding the peak intensity for each spectrum at $\approx\qty{665}{\nm}$. The thus determined peak intensity was plotted vs time in the bottom-right inset, \cref{fig-a:MB-N04H-small-nostir-photodegradation}. The decolourisation proceeded quickly and reached a minimum after \qty{25}{\minute}, then rebounded and continued at a steady rate until the end of the experiment.
Tauc analysis of the optical band gap from the absorption data of the \ch{ZnO} quantum dots with a larger starting size: \qty{5.8}{\nm}, \qty{6.7}{\nm}, and \qty{7.4}{\nm}, respectively are presented in condensed form in \cref{fig:MB-panels-nostir-photodegradation}. Detailed plots along the lines of the first plot for these experiments are available in the SI.
For starting diameters of \qty{5.8}{\nm} (as well as starting sizes of \qty{3.2}{\nm}), a complete decolourisation is seen during the experiment while the particles with initially larger sizes (\qty{6.7}{\nm} and \qty{7.4}{\nm}) show faster initial kinetics but also a marked minimum after around \qty{25}{\minute}, then rebounding and continuing to proceed at a steady rate until the end of the experiment. The reason behind this is not clear but was seen for repeated experiments and is related to \begin{enumerate*}[label=\alph*)]%
\item faster initial decolourisation kinetics in combination with
\item a slow secondary decolourisation process extending significantly beyond \qty{150}{\minute} (\cref{fig:MB-N02A-small-nostir-mbabs,fig:MB-N04H-medium-nostir-mbabs}).
\end{enumerate*}
One possible explanation is that the larger particles to a larger extent caused reduction of methylene blue and thus deactivated the oxidation process. To fully resolve this, spectral signatures and mass spectrosopy is needed and is an interesting topic for a more specialised study of possibly different mechanisms dominating for different sizes and extent of exposed Miller indices of the ZnO particles. The results show that a considerable increase in the photocatalytic activity can be achieved by starting with low-dimensional \ch{ZnO} ($<\qty{6}{\nm}$) where the different behaviour for the larger sizes instead can be utilized in experiments where size selective effects are desired.
\begin{figure*}[tbp]
\centering
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=6.80in]{figure/fig-MB-N04H-small-nostir-photodegradation-1}
}
\end{knitrout}
{\phantomsubcaption\label{fig-a:MB-N04H-small-nostir-photodegradation}}
{\phantomsubcaption\label{fig-b:MB-N04H-small-nostir-photodegradation}}
\caption{Optical absorption spectra recorded under 1~Sun illumination from $t=0$ (when ultra-small \ch{ZnO(EtOH)} nanoparticles were added to the \ch{MB(aq)} dye solution) until $t>\qty{2}{\hour}$ with $\Delta t=\qty{1}{\minute}$.
The top inset (\subref{fig-a:MB-N04H-small-nostir-photodegradation}) shows the fitted height of the strongest methylene blue absorption peak vs time.
The bottom inset (\subref{fig-b:MB-N04H-small-nostir-photodegradation}) shows the result of a Tauc analysis on the band edge (the Tauc fits are highlighted by the yellow lines intersecting the band edge region of each spectrum in the main plot).}
\label{fig:MB-N04H-small-nostir-photodegradation}
\end{figure*}
\begin{figure*}[tbp]
\centering
\begin{subfigure}[b]{0.33\linewidth}
\caption{}%a
\label{fig:MB-N02A-small-nostir-mbabs}
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=2.30in]{figure/fig-MB-N02A-small-nostir-mbabs-1}
}
\end{knitrout}
\end{subfigure}%
\hfill%
\begin{subfigure}[b]{0.33\linewidth}
\caption{}%b
\label{fig:MB-N04H-medium-nostir-mbabs}
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=2.30in]{figure/fig-MB-N04H-medium-nostir-mbabs-1}
}
\end{knitrout}
\end{subfigure}%
\hfill%
\begin{subfigure}[b]{0.33\linewidth}
\caption{}%c
\label{fig:MB-N02A-large-nostir-mbabs}
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=2.30in]{figure/fig-MB-N02A-large-nostir-mbabs-1}
}
\end{knitrout}
\end{subfigure}\\%
\begin{subfigure}[b]{0.33\linewidth}
\caption{}%d
\label{fig:MB-N02A-small-nostir-edge}
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=2.30in]{figure/fig-MB-N02A-small-nostir-edge-1}
}
\end{knitrout}
\end{subfigure}%
\hfill%
\begin{subfigure}[b]{0.33\linewidth}
\caption{}%e
\label{fig:MB-N04H-medium-nostir-edge}
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=2.30in]{figure/fig-MB-N04H-medium-nostir-edge-1}
}
\end{knitrout}
\end{subfigure}%
\hfill%
\begin{subfigure}[b]{0.33\linewidth}
\caption{}%f
\label{fig:MB-N02A-large-nostir-edge}
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=2.30in]{figure/fig-MB-N02A-large-nostir-edge-1}
}
\end{knitrout}
\end{subfigure}
%
\caption{Absorption of the dye during the decolourisation process, and band gap and size of the growing ZnO QDs for particles with starting diameters of
(\subref{fig:MB-N02A-small-nostir-mbabs}, \subref{fig:MB-N02A-small-nostir-edge}) \qty{3.2}{\nm},
(\subref{fig:MB-N04H-medium-nostir-mbabs}, \subref{fig:MB-N04H-medium-nostir-edge}) \qty{5.8}{\nm}, and
(\subref{fig:MB-N02A-large-nostir-mbabs}, \subref{fig:MB-N02A-large-nostir-edge}) \qty{6.7}{\nm}.}
\label{fig:MB-panels-nostir-photodegradation}
\end{figure*}
\begin{figure*}[tbp]
\centering
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=6.80in]{figure/fig-MB-N04E-large-nostir-photodegradation-1}
}
\end{knitrout}
{\phantomsubcaption\label{fig-a:MB-N04E-large-nostir-photodegradation}}
{\phantomsubcaption\label{fig-b:MB-N04E-large-nostir-photodegradation}}
\caption{Optical absorption spectra recorded under 1~Sun illumination from $t=0$ until $t>\qty{2}{\hour}$ with $\Delta t=\qty{1}{\minute}$.
The cuvette contained methylene blue in \ch{H2O}:EtOH 90:10 (by volume) in the presence of \ch{ZnO} nanoparticles with $d<\qty{10}{\nm}$ in suspension.
The top inset (\subref{fig-a:MB-N04E-large-nostir-photodegradation}) shows the fitted height of the strongest methylene blue absorption peak vs time.
The bottom inset (\subref{fig-b:MB-N04E-large-nostir-photodegradation}) shows the result of a Tauc analysis on the band edge (the Tauc fits are highlighted by the yellow lines intersecting the band edge region of each spectrum in the main plot).}
\label{fig:MB-N04E-large-nostir-photodegradation}
\end{figure*}
%%% MB data: kinetics
\begin{figure*}[tbp]
\centering
\begin{knitrout}\footnotesize
\definecolor{shadecolor}{rgb}{0.969, 0.969, 0.969}\color{fgcolor}
{\centering \includegraphics[width=6.80in]{figure/fig-MB-ratio-fityk-CC0-1}
}
\end{knitrout}
\caption{Decolourisation of MB with \ch{ZnO} NPs of varying initial diameter, $d_\text{i}$.}
\label{fig:MB-ratio-fityk-CC0}
\end{figure*}
\begin{table}[tbp]
\centering
\begin{small}
\caption{Fitted rate constants for the two different regions of the decolourisation curves, per \cref{eq:logconc-kobs}.}
\label{tab:kinpar-MB}
% latex table generated in R 4.0.3 by xtable 1.8-4 package
% Fri Aug 6 06:01:55 2021
\begin{tabular}{S[table-format=1.2]S[table-format=1.2]S[table-format=1.2]S[table-format=1.2]}
\toprule
& \multicolumn{3}{c}{Rate constant, $k_\text{obs}$/\num[retain-unity-mantissa = false]{1E-4}\unit{\per\second}} \\ \cmidrule(lr){2-4}
{$d_\text{i}/\unit{\nm}$} & {$t<\qty{20}{\min}$} & {$t>\qty{35}{\min}$} & {Averaged} \\
\midrule
3.18 & 0.96 & 6.55 & 3.76 \\
3.23 & 1.08 & 0.59 & 0.83 \\
5.85 & 9.27 & 0.56 & 4.91 \\
6.72 & 9.94 & 0.28 & 5.11 \\
7.42 & 14.47 & 0.56 & 7.52 \\
\bottomrule
\end{tabular}
\end{small}
\end{table}
Apart from the rate of electrons and holes reaching the surface of the \ch{ZnO} particles and the reactivation of the active sites after the first reaction, the decolourisation kinetics also depends on the transport of reactants, by molecular diffusion and/or convective transport. This in turn can be changed by adding a velocity gradient (flow near surface) or a changed concentration gradient. In this study we do not aim for an optimised degradation speed (where stirring or higher concentration of the model pollutant would be beneficial) but instead aim to observe the inherent kinetics of the process. In either case, the surface reactions create a surface concentration different from the bulk liquid and allow a concentration difference to exist between the bulk and the reactive surface. This difference is also related to challenges in correlating the intrinsic reaction rate, $r$, using the measurable bulk concentration $C_\text{b}$ rather than the unattainable surface concentration, $C_\text{s}$. Nevertheless, utilising a system without external stirring and at ambient conditions would give a relative comparison of the kinetics of decolourisation between the ZnO quantum dots growing from different starting sizes.
Under steady state conditions, intermediates such as \ch{^.OH}, \ch{e^-(aq)}, and \ch{O2^{.-}} that destroy the dye molecules and eventually mineralise the pollutant, may in an ideal situation react exclusively with the intact dye molecules with decolourisation as a result. A rate equation for dye decolourisation can then be given by the differential equation:
\begin{equation}
-\frac{d[\text{dye}]}{dt} = k_\text{eff} [\text{int}] [\text{dye}]