Processing math: 100%
Research article Topical Sections

Impact of particle shape on electron transport and lifetime in zinc oxide nanorod-based dye-sensitized solar cells

  • Owing to its high electron mobility, zinc oxide represents a promising alternative to titanium dioxide as the working electrode material in dye-sensitized solar cells (DSCs). When zinc oxide is grown into 1-D nanowire arrays and incorporated into the working electrode of DSCs, enhanced electron dynamics and even a decoupling of electron transport (τd) and electron lifetime (τn) have been observed. In this work, DSCs with working electrodes composed of solution-grown, unarrayed ZnO nanorods are investigated. In order to determine whether such devices give rise to similar decoupling, intensity modulated photocurrent and photovoltage spectroscopies are used to measure τd and τn, while varying the illumination intensity. In addition, ZnO nanorod-based DSCs are compared with ZnO nanoparticle-based DSCs and nanomaterial shape is shown to affect electron dynamics. Nanorod-based DSCs exhibit shorter electron transport times, longer electron lifetimes, and a higher τnd ratio than nanoparticle-based DSCs.

    Citation: Roger Chang, Kemakorn Ithisuphalap, Ilona Kretzschmar. Impact of particle shape on electron transport and lifetime in zinc oxide nanorod-based dye-sensitized solar cells[J]. AIMS Materials Science, 2016, 3(1): 51-65. doi: 10.3934/matersci.2016.1.51

    Related Papers:

    [1] Iyad Suwan, Mohammed S. Abdo, Thabet Abdeljawad, Mohammed M. Matar, Abdellatif Boutiara, Mohammed A. Almalahi . Existence theorems for $ \Psi $-fractional hybrid systems with periodic boundary conditions. AIMS Mathematics, 2022, 7(1): 171-186. doi: 10.3934/math.2022010
    [2] Iman Ben Othmane, Lamine Nisse, Thabet Abdeljawad . On Cauchy-type problems with weighted R-L fractional derivatives of a function with respect to another function and comparison theorems. AIMS Mathematics, 2024, 9(6): 14106-14129. doi: 10.3934/math.2024686
    [3] Chutarat Treanbucha, Weerawat Sudsutad . Stability analysis of boundary value problems for Caputo proportional fractional derivative of a function with respect to another function via impulsive Langevin equation. AIMS Mathematics, 2021, 6(7): 6647-6686. doi: 10.3934/math.2021391
    [4] Abdelatif Boutiara, Mohammed S. Abdo, Manar A. Alqudah, Thabet Abdeljawad . On a class of Langevin equations in the frame of Caputo function-dependent-kernel fractional derivatives with antiperiodic boundary conditions. AIMS Mathematics, 2021, 6(6): 5518-5534. doi: 10.3934/math.2021327
    [5] Songkran Pleumpreedaporn, Chanidaporn Pleumpreedaporn, Weerawat Sudsutad, Jutarat Kongson, Chatthai Thaiprayoon, Jehad Alzabut . On a novel impulsive boundary value pantograph problem under Caputo proportional fractional derivative operator with respect to another function. AIMS Mathematics, 2022, 7(5): 7817-7846. doi: 10.3934/math.2022438
    [6] Zaid Laadjal, Fahd Jarad . On a Langevin equation involving Caputo fractional proportional derivatives with respect to another function. AIMS Mathematics, 2022, 7(1): 1273-1292. doi: 10.3934/math.2022075
    [7] Yujun Cui, Chunyu Liang, Yumei Zou . Existence and uniqueness of solutions for a class of fractional differential equation with lower-order derivative dependence. AIMS Mathematics, 2025, 10(2): 3797-3818. doi: 10.3934/math.2025176
    [8] Asghar Ahmadkhanlu, Hojjat Afshari, Jehad Alzabut . A new fixed point approach for solutions of a $ p $-Laplacian fractional $ q $-difference boundary value problem with an integral boundary condition. AIMS Mathematics, 2024, 9(9): 23770-23785. doi: 10.3934/math.20241155
    [9] Sabri T. M. Thabet, Miguel Vivas-Cortez, Imed Kedim . Analytical study of $ \mathcal{ABC} $-fractional pantograph implicit differential equation with respect to another function. AIMS Mathematics, 2023, 8(10): 23635-23654. doi: 10.3934/math.20231202
    [10] Özge Arıbaş, İsmet Gölgeleyen, Mustafa Yıldız . On the solvability of an initial-boundary value problem for a non-linear fractional diffusion equation. AIMS Mathematics, 2023, 8(3): 5432-5444. doi: 10.3934/math.2023273
  • Owing to its high electron mobility, zinc oxide represents a promising alternative to titanium dioxide as the working electrode material in dye-sensitized solar cells (DSCs). When zinc oxide is grown into 1-D nanowire arrays and incorporated into the working electrode of DSCs, enhanced electron dynamics and even a decoupling of electron transport (τd) and electron lifetime (τn) have been observed. In this work, DSCs with working electrodes composed of solution-grown, unarrayed ZnO nanorods are investigated. In order to determine whether such devices give rise to similar decoupling, intensity modulated photocurrent and photovoltage spectroscopies are used to measure τd and τn, while varying the illumination intensity. In addition, ZnO nanorod-based DSCs are compared with ZnO nanoparticle-based DSCs and nanomaterial shape is shown to affect electron dynamics. Nanorod-based DSCs exhibit shorter electron transport times, longer electron lifetimes, and a higher τnd ratio than nanoparticle-based DSCs.


    In this work, we are concerned with the following fractional boundary value problem

    $ Dα, φ0+u(t)+λf(t,u(t))=0,   0<t<1,
    $
    (1.1)
    $ u(0)=0,   u(1)=m2i=1βiu(ηi),
    $
    (1.2)

    where $ 1 < \alpha\leq 2, \ D^{\alpha, \ \varphi}_{0^{+}} $ is the fractional derivative operator of order $ \alpha $ with respect to a certain strictly increasing function $ \varphi\in C^{2}[0, 1] $ with $ \varphi'(x) > 0 $ for all $ x\in [0, 1] $ and $ f, \varphi, \beta_{i}, \eta_{i}, \lambda $ satisfy:

    $ (H_{1})\ f(t, u):[0, 1]\times R^{+}\rightarrow R^{+} $ is continuous and increasing in $ u $ for each $ t\in [0, 1], $ the function $ \varphi:[0, 1]\rightarrow R $ is a strictly increasing function such that $ \varphi\in C^{2}[0, 1], $ with $ \varphi'(x) > 0 $ for all $ x\in [0, 1] $;

    $ (H_{2})\ 0\leq \beta_{i} < 1, \ 0 < \eta_{i} < 1\ (i = 1, 2, \cdot\cdot\cdot, m-2) $ satisfy $ 0\leq \sum\limits_{i = 1}^{m-2}\beta_{i} < 1, \ 0 < \eta_{1} < \eta_{2} < \cdot\cdot\cdot < \eta_{m-2} < 1, \lambda > 0 $ is a parameter.

    The fractional calculus first began in 1965. After a long time, this subject was relevant only in pure mathematics, many mathematicians have studied these new fractional operators by presenting new definitions and studying their important properties, but in the last century, this subject showed its applications in the modeling of many phenomena in various fields of science and engineering, such as control theory, electric circuits, viscoelasticity, mechanics, physics, neural networks [1,2,3,4,5,6,7,8,9,10,11,12].

    In all those definitions there is a special kind of a kernel dependency. Therefore, in order to analyze fractional differential equations in a generic way, a fractional derivative with respect to another function called $ \varphi $-Riemann-Liouville and $ \varphi $-Caputo derivative was proposed [15] and for particular choices of $ \varphi, $ we obtain some well known fractional derivative. For example, if we choose $ \varphi(t) = t, $ the $ \varphi $-Riemann-Liouville and $ \varphi $-Caputo fractional derivative are reduced to the Riemann-Liouville and Caputo derivative respectively in traditional sense. Alternatively, if we choose $ \varphi(t) = logt, $ where $ log(\cdot) = log_{e}(\cdot), $ the $ \varphi $-Riemann-Liouville and $ \varphi $-Caputo fractional derivative are reduced to the Hadamard and Hadamard-Caputo derivative respectively [13,14]. Therefore, problem (1.1), (1.2) generates many types and also mixed types of fractional differential equations with boundary conditions.

    Let us mention some motivations for studying problem (1.1), (1.2). In the limit case $ \lambda = 1 $ and $ \beta_{i} = \beta, \eta_{i} = \eta, i = 1, 2, \cdot\cdot\cdot, $ (1.1), (1.2) reduces to

    $ Dα, φ0+u(t)+f(t,u(t))=0,   0<t<1,
    $
    (1.3)
    $ u(0)=0,   u(1)=βu(η).
    $
    (1.4)

    In [22], the authors studied the existence of positive solutions for the problem (1.3), (1.4). In [23], using the fixed point theorems of Banach and Schaefer, the authors studied the existence and uniqueness results of the boundary value problems for a fractional differential equation involving Caputo operator with respect to the new function $ \psi $ given by the form

    $ ^{C}D_{a^{+}}^{\alpha, \ \psi}y(t) = f(t, y(t)), \ \ \ \ t\in [a, b], $
    $ y_{\psi}^{[k]}(a) = y_{a}^{k}, \ \ \ k = 0, 1, \cdot\cdot\cdot, n-2;\ \ \ y_{\psi}^{[n-1]}(b) = y_{b}, $

    where $ ^{C}D_{a^{+}}^{\alpha, \ \psi} $ is the $ \psi $-Caputo fractional derivative of order $ n-1 < \alpha\leq n\ (n = [\alpha]+1). $

    Almeida et al. [24] studied the initial value problem of a fractional differential equation including $ \varphi $-Caputo fractional derivative

    $ ^{C}D_{a^{+}}^{\alpha, \ \varphi}y(t) = f(t, y(t)), \ \ \ \ \ t\in [a, b], $
    $ y(a) = y_{a}, \ \ \ y_{\psi}^{[k]}(a) = y_{a}^{k}, \ \ \ k = 1, 2, \cdot\cdot\cdot, n-1 $

    where $ ^{C}D_{a^{+}}^{\alpha, \ \varphi} $ is the $ \varphi $-Caputo fractional derivative of order $ n-1 < \alpha < n. $

    On the other hand, most of the existing literature does not consider the properties of the solution. Well, here's the problem: How can we get the properties of the solutions when they are known to exist. So far, few papers can be found in the literature both on the existence and properties of solutions for fractional boundary value problem.

    Zhai et al. [16] studied the following nonlinear fractional four-point boundary value problem with a parameter:

    $ D_{0^{+}}^{\alpha}u(t)+\lambda f(t, u(t)) = 0, \ \ 0 \lt t \lt 1, $
    $ u'(0)-\mu_{1}u(\xi) = 0, \ \ u'(1)+\mu_{2}u(\eta) = 0, $

    where $ D_{0^{+}}^{\alpha} $ is the Riemann-Liouville fractional derivative. Unfortunately, the fractional derivative in this paper is just the Caputo derivative in the traditional sense.

    Inspired by the work in [16] and many known results, in this paper, we develop the existence and uniqueness of positive solutions to the fractional boundary value problems (1.1), (1.2). Using the fixed point theorems of concave operators in partial ordering Banach spaces, we not only show the existence of positive solution for the problem (1.1), (1.2), but also present some properties of positive solutions to the boundary value problem dependent on the parameter. Some definitions, lemmas and basic results about $ \varphi $-Riemann-Liouville derivative are presented in Section 2. The main results are given in section 3. Some illustrative examples are constructed in Section 4.

    Let $ \varphi $ be a function of $ C^{2}[a, b] $ with $ \varphi'(t) > 0 $ for all $ t\in [a, b]. $

    Definition 2.1 [21] The $ \varphi $-Riemann-Liouville fractional integral of order $ \alpha > 0 $ of a function $ x:[a, b]\rightarrow R $ is given by

    $ I^{\alpha, \ \varphi}_{a^{+}}x(t): = \frac{1}{\Gamma (\alpha)}\int_{a}^{t}\varphi'(s)(\varphi(t)-\varphi(s))^{\alpha-1}x(s)ds. $

    Definition 2.2 [21] The $ \varphi $-Riemann-Liouville fractional derivative of order $ \alpha $ of a function $ x $ is defined as

    $ D^{\alpha, \ \varphi}_{a^{+}}x(t): = \biggl( \frac{1}{\varphi'(t)}\frac{d}{dt}\biggr)^{n}I^{n-\alpha, \ \varphi}_{a^{+}}x(t) = \frac{1}{\Gamma (n-\alpha)}\biggl(\frac{1}{\varphi'(t)}\frac{d}{dt}\biggr)^{n}\int_{a}^{t}\varphi'(s)(\varphi(t)-\varphi(s))^{n-\alpha-1}x(s)ds, $

    where $ n = [\alpha]+1. $

    The $ \varphi $-Caputo fractional derivative of order $ \alpha $ of a function $ x $ is defined as

    $ ^{C}D^{\alpha, \ \varphi}_{a^{+}}x(t): = D^{\alpha, \ \varphi}_{a^{+}}\biggl[x(t)-\sum\limits_{k = 0}^{n-1} \frac{x_{\varphi}^{[k]}(a)}{k!} (\varphi(t)-\varphi(a))^{k}\biggr], \ n = [\alpha]+1\ \mbox{for}\ \alpha \not\in N, \ n = \alpha\ \mbox{for} \ \alpha \in N, $

    where $ x_{\varphi}^{[k]}(t): = \biggl(\frac{1}{\varphi'(t)}\frac{d}{dt}\bigamma)^{k}x(t). $

    Theorem 2.1 [15] If $ x\in C^{n}[a, b], $ then

    $ ^{C}D^{\alpha, \ \varphi}_{a^{+}}x(t) = I^{n-\alpha, \ \varphi}_{a^{+}}\biggl( \frac{1}{\varphi'(t)}\frac{d}{dt}\biggr)^{n}x(t). $

    Theorem 2.2 [15] Let $ x: [a, b]\rightarrow R. $ Then

    1. If $ x\in C[a, b], $ then $ ^{C}D^{\alpha, \ \varphi}_{a^{+}}I^{\alpha, \ \varphi}_{a^{+}}x(t) = x(t); $

    2. If $ x\in C^{n-1}[a, b], $ then

    $ I^{\alpha, \ \varphi\ C}_{a^{+}} D^{\alpha, \ \varphi}_{a^{+}}x(t) = x(t)-\sum\limits_{k = 0}^{n-1} \frac{x_{\varphi}^{[k]}(a)}{k!} (\varphi(t)-\varphi(a))^{k}. $

    In the following, we give some basic definitions in ordered Banach spaces and two fixed point theorems of concave operators which will be used later, all the materials can be found in [17,18,19,20]. We denote $ \theta $ the zero element of $ E. $

    Definition 2.3 Let $ X $ be a semi-ordered real Banach space. The nonempty convex closed subset $ P $ of $ X $ is called a cone in $ X $ if $ ax\in P $ for all $ x\in P $ and $ a\geq 0 $ and $ x\in X $ and $ -x\in X $ imply $ x = \theta. $

    Let the real Banach space $ E $ be partially ordered by a cone $ P $ of $ E, $ that is $ x\leq y $ if and only if $ y-x\in P. $ If $ x\leq y $ and $ x\neq y, $ then we denote $ x < y $ or $ y > x. $

    Recall that cone $ P $ is said to be solid if the interior $ \mathring{P} $ is nonempty and we denote $ x > > 0 $ if $ x\in \mathring{P}. $ $ P $ is normal if there exists a positive constant $ N $ such that $ \theta\leq x\leq y $ implies $ \|x\|\leq N\|y\|, $ $ N $ is called the normal constant of $ P. $

    Definition 2.4 Let $ D $ be a convex subset in $ E. $ An operator $ T:D\rightarrow E $ is a concave operator if

    $ T(tx+(1-t)y)\geq tTx+(1-t)Ty $

    for all $ x, y\in D $ and $ t\in [0, 1]. $

    Definition 2.5 Let $ P $ be a cone in real Banach space $ E $ and $ e\in P\backslash\{\theta\}. $ Set

    $ E_{e} = \{x\in E:\mbox{there exists}\ \lambda \gt 0\ \mbox{such that}-\lambda e\leq x\leq \lambda e\} $

    and define

    $ \|x\|_{e} = inf\{\lambda \gt 0:-\lambda e\leq x\leq \lambda e\}, \ \ \ \forall x\in E_{e}. $

    It is easy to see that $ E_{e} $ becomes a normed linear space under the norm $ \|\cdot\|_{e}. $ $ \|x\|_{e} $ is called the e-norm of the element $ x\in E_{e}. $

    Lemma 2.3 [17] Suppose $ P $ is a normal cone. Then the following results hold:

    (i) $ E_{e} $ is a Banach space with e-norm, and there exists a constant $ m > 0 $ such that $ \|x\|\leq m\|x\|_{e}, \ \ \ \forall x\in E_{e}; $

    (ii) $ P_{e} = E_{e}\cap P $ is a normal solid cone of $ E_{e} $ and $ \mathring{P_{e}} = \{x\in E_{e}:\mbox{there exists}\ \delta = \delta(x) > 0\ \mbox{such that} \ x\geq \delta e\}. $

    Lemma 2.4 [19] Suppose $ P $ is a normal solid cone and operator $ T:P\rightarrow P $ is a concave operator. Assume that $ T\theta > > \theta. $ Then the following results hold:

    (ⅰ) there exists $ 0 < \lambda^{*}\leq \infty $ such that the equation

    $ u=λTu
    $
    (2.1)

    has a unique solution $ u(\lambda) $ in $ P $ for $ 0\leq \lambda < \lambda^{*}, $ when $ \lambda\geq \lambda^{*}, $ the operator equation (2.1) has no solution in $ P; $

    (ⅱ) for any $ u_{0}\in P, $ let $ u_{0}(\lambda) = u_{0}, \ \ u_{n}(\lambda) = \lambda Tu_{n-1}(\lambda), \ n = 1, 2, 3, \cdot\cdot\cdot. $ Then for $ 0 < \lambda < \lambda^{*}, $ we have $ u_{n}(\lambda)\rightarrow u(\lambda) $ as $ n\rightarrow \infty; $

    (ⅲ) $ u(\cdot):[0, \lambda^{*})\rightarrow P $ is continuous and strongly increasing $ (0\leq \lambda_{1} < \lambda_{2} < \lambda^{*}\Rightarrow u(\lambda_{1}) < < u(\lambda_{2})). $ Furthermore, $ u(t\lambda)\leq tu(\lambda) $ for $ 0\leq \lambda < \lambda^{*}, \ 0\leq t\leq 1. $

    (ⅳ) $ \lambda\rightarrow\lambda^{*}-0 $ implies $ \|u(\lambda)\|\rightarrow \infty; $

    (ⅴ) if there exist $ v_{0}\in P $ and $ \lambda_{0} > 0 $ such that $ \lambda_{0}Tv_{0}\leq v_{0}, $ then $ \lambda^{*} > \lambda_{0}. $

    Lemma 2.5 [19] If $ T:P\rightarrow P $ is concave, then $ T $ is increasing.

    Lemma 2.6 [20] Suppose $ P $ is a normal solid cone and operator $ T:\mathring{P}\rightarrow \mathring{P} $ is an increasing operator. We presume there exists a constant $ 0 < r < 1 $ such that

    $ T(tu)\geq t^{r}Tu, \ \ \ \forall u\in \mathring{P}, \ 0 \lt t \lt 1. $

    Let $ u_{\lambda} $ be the unique solution in $ \mathring{P} $ of the equation $ Tu = \lambda u(\lambda > 0). $ Then

    (ⅰ) $ u_{\lambda} $ is strongly decreasing $ (0 < \lambda_{1} < \lambda_{2}\Rightarrow u_{\lambda_{1}} > > u_{\lambda_{2}}); $

    (ⅱ) $ u_{\lambda} $ is continuous $ (\lambda\rightarrow \lambda_{0}(\lambda_{0} > 0)\Rightarrow \|u_{\lambda}-u_{\lambda_{0}}\|\rightarrow 0); $

    (ⅲ) $ \lim_{\lambda\rightarrow +\infty}\|u_{\lambda}\| = 0, \ \ \ \lim_{\lambda\rightarrow 0^{+}}\|u_{\lambda}\| = +\infty. $

    Denote

    $ \mu = (\varphi(1)-\varphi(0))^{\alpha-1}-\sum\limits_{i = 1}^{m-2}\beta_{i}(\varphi(\eta_{i})-\varphi(0))^{\alpha-1}. $

    Lemma 3.1 Assume that $ (H_{1}), (H_{2}) $ hold. Then the fractional boundary value problem (1.1), (1.2) is equivalent to the following integral equation

    $ u(t)=10G(t,s)φ(s)λf(s,u(s))ds,
    $
    (3.1)

    where

    $ G(t,s)={m2j=1βj(φ(ηj)φ(0))α1(φ(t)φ(s))α1(φ(t)φ(s))α1(φ(1)φ(0))α1μΓ(α)+(φ(t)φ(0))α1(φ(1)φ(s))α1m2j=iβj(φ(ηj)φ(s))α1(φ(t)φ(0))α1μΓ(α)0t1, ηi1smin{ηi,t},i=1,2,,m1;(φ(t)φ(0))α1(φ(1)φ(s))α1m2j=iβj(φ(ηj)φ(s))α1(φ(t)φ(0))α1μΓ(α)0t1, max{ηi1,t}sηi, i=1,2,,m1.
    $
    (3.2)

    Here for the sake of convenience, we write $ \eta_{0} = 0, \eta_{m-1} = 1 $ and $ \sum\limits_{i = m_{1}}^{m_{2}}f_{i} = 0 $ for $ m_{2} < m_{1}. $

    Proof. Use $ y(t) $ to replace $ f(t, u) $ in (1.1). Let

    $ D^{\alpha, \ \varphi}_{0^{+}}u(t)+\lambda y(t) = 0. $

    Then from Theorem 2.2, we have

    $ u(t)=c1(φ(t)φ(0))α1+c2(φ(t)φ(0))α21Γ(α)t0φ(s)(φ(t)φ(s))α1λy(s)ds.
    $
    (3.3)

    Note that $ u(0) = 0 $ implies $ c_{2} = 0. $ Therefore, we obtain

    $ u(t)=c1(φ(t)φ(0))α11Γ(α)t0φ(s)(φ(t)φ(s))α1λy(s)ds.
    $
    (3.4)

    In particular,

    $ u(1) = c_{1}(\varphi(1)-\varphi(0))^{\alpha-1} -\frac{1}{\Gamma(\alpha)}\int_{0}^{1}\varphi'(s)(\varphi(1)-\varphi(s))^{\alpha-1}\lambda y(s)ds $

    and

    $ u(\eta_{i}) = c_{1}(\varphi(\eta_{i})-\varphi(0))^{\alpha-1} -\frac{1}{\Gamma(\alpha)}\int_{0}^{\eta_{i}}\varphi'(s)(\varphi(\eta_{i})-\varphi(s))^{\alpha-1}\lambda y(s)ds. $

    Note that $ u(1) = \sum_{i = 1}^{m-2}\beta_{i}u(\eta_{i}), $ and hence we obtain

    $ c_{1} = \frac{1}{\mu\Gamma(\alpha)}\int_{0}^{1}\varphi'(s)(\varphi(1)-\varphi(s))^{\alpha-1}\lambda y(s)ds-\frac{\sum\limits_{i = 1}^{m-2}\beta_{i}}{\mu\Gamma(\alpha)}\int_{0}^{\eta_{i}}\varphi'(s)(\varphi(\eta_{i})-\varphi(s))^{\alpha-1}\lambda y(s)ds. $

    As a result, from (3.4) we have

    $ u(t)=1Γ(α)t0φ(s)(φ(t)φ(s))α1λy(s)ds+(φ(t)φ(0))α1μΓ(α)10φ(s)(φ(1)φ(s))α1λy(s)dsm2i=1βi(φ(t)φ(0))α1μΓ(α)ηi0φ(s)(φ(ηi)φ(s))α1λy(s)ds=10G(t,s)φ(s)λy(s)ds.
    $
    (3.5)

    This completes the proof. Lemma 3.2 Assume that $ (H_{1}), (H_{2}) $ hold. Then the Green's function $ G(t, s) $ defined by (3.2) has the following properties:

    (ⅰ) $ G(t, s) $ is a continuous function on $ [0, 1]\times [0, 1]; $

    (ⅱ) $ G(t, s) > 0 $ for all $ (t, s)\in (0, 1); $

    (ⅲ)

    $ G(t,s)(φ(1)φ(0))α1(φ(1)φ(s))α1μΓ(α),  s(0,1).
    $
    (3.6)

    Proof. (ⅰ) From the expression of $ G(t, s), $ it is easy to get that $ G(t, s) $ is continuous.

    (ⅱ) From $ (H_{1}), (H_{2}), $ we can get $ \mu > 0. $

    For $ 0\leq t \leq 1, \ \eta_{i-1}\leq s\leq \min\{\eta_{i}, t\}, i = 1, 2, \cdot\cdot\cdot, m-1, $

    $ G(t,s)=1μΓ(α)[(φ(t)φ(0))α1(φ(1)φ(s))α1m2j=iβj(φ(ηj)φ(s))α1(φ(t)φ(0))α1μ(φ(t)φ(s))α1]=(φ(t)φ(0))α1μΓ(α)[(φ(1)φ(s))α1m2j=iβj(φ(ηj)φ(s))α1μ(φ(t)φ(s)φ(t)φ(0))α1].
    $

    Consider

    $ e(t) = (\varphi(1)-\varphi(s))^{\alpha-1} -\sum\limits_{j = i}^{m-2}\beta_{j}(\varphi(\eta_{j})-\varphi(s))^{\alpha-1} -\mu \biggl( \frac{\varphi(t)-\varphi(s)}{\varphi(t)-\varphi(0)}\biggr)^{\alpha-1}. $

    Then we have

    $ e'(t) = -\mu(\alpha-1) \biggl( \frac{\varphi(t)-\varphi(s)}{\varphi(t)-\varphi(0)}\biggr)^{\alpha-2} \frac{\varphi'(t)(\varphi(s)-\varphi(0))}{(\varphi(t)-\varphi(0))^{2}}\leq 0, $

    which implies that $ e(t) $ is decreasing for $ 0\leq t \leq 1, \ \eta_{i-1}\leq s\leq \min\{\eta_{i}, t\}, i = 1, 2, \cdot\cdot\cdot, m-1. $

    Furthermore, we have

    $ e(1)=(φ(1)φ(s))α1m2j=iβj(φ(ηj)φ(s))α1μ(φ(1)φ(s)φ(1)φ(0))α1=m2j=1βj(φ(ηj)φ(0))α1(φ(1)φ(s))α1(φ(1)φ(0))α1m2j=iβj(φ(ηj)φ(s))α1=m2j=1βj(φ(ηj)φ(0))α1(φ(1)φ(s))α1(φ(1)φ(0))α1m2j=iβj(φ(ηj)φ(0))α1(φ(ηj)φ(s))α1(φ(ηj)φ(0))α1.
    $

    As we know,

    $ \frac{(\varphi(1)-\varphi(s))^{\alpha-1}} {(\varphi(1)-\varphi(0))^{\alpha-1}} \gt \frac{(\varphi(\eta_{j})-\varphi(s))^{\alpha-1}} {(\varphi(\eta_{j})-\varphi(0))^{\alpha-1}}, $

    consequently, we have $ e(1) > 0. $

    As a result, $ G(t, s) > 0 $ for $ 0\leq t \leq 1, \ \eta_{i-1}\leq s\leq \min\{\eta_{i}, t\}, i = 1, 2, \cdot\cdot\cdot, m-1. $

    For $ 0\leq t \leq 1, \ \max\{\eta_{i-1}, t\}\leq s\leq \eta_{i}, i = 1, 2, \cdot\cdot\cdot, m-1, $

    $ G(t,s)=1μΓ(α)[(φ(t)φ(0))α1(φ(1)φ(s))α1m2j=iβj(φ(ηj)φ(s))α1(φ(t)φ(0))α1]=1μΓ(α)(φ(t)φ(0))α1[(φ(1)φ(s))α1m2j=iβj(φ(ηj)φ(s))α1]>0.
    $

    Therefore, $ G(t, s) > 0 $ for all $ t, s\in (0, 1). $

    (ⅲ) From $ (H_{1}), (H_{2}) $ and the expression of $ G(t, s), $ we can easily obtain this property. Let $ E = C[0, 1] $ be a Banach space equipped with the norm

    $ \|u\|_{E} = \max\limits_{t\in [0, 1]}|u(t)|. $

    Define the cone $ K\subset E $ by

    $ K = \{u\in E:u(t)\geq 0, \ 0\leq t\leq 1\}, $

    then $ K $ is a normal solid cone of $ E. $ Let

    $ g(t)=10G(t,s)φ(s)ds,   0t1.
    $
    (3.7)

    From $ (H_{1}) $ and Lemma 3.2, we have $ g(t)\in E $ and $ g(t)\in K\setminus\{\theta\}. $

    Let $ X = E_{e} = \{u\in E:\exists \tau > 0, \ \mbox{such that}\ -\tau g(t)\leq u(t)\leq \tau g(t), \ 0\leq t\leq 1\} $ be a Banach space equipped with the norm $ \|u\|_{X} = \inf\{\tau > 0: -\tau g(t)\leq u(t)\leq \tau g(t), \ \ 0\leq t\leq 1\}. $

    Define $ \tilde{K} = X\cap K, $ then $ \tilde{K} $ is a normal solid cone of $ X $ and $ \mathring{\tilde{K}} = \{u\in X: \mbox{there exists}\ \delta > 0 \ \mbox{such that}\ u(t)\geq \delta g(t), \ \ 0\leq t\leq 1, \ \ \forall u\in X\}. $ Moreover, there exists a constant $ m > 0 $ such that $ \|u\|_{E}\leq m \|u\|_{X}. $

    Theorem 3.3 Assume that $ (H_{1}), (H_{2}) $ hold. Furthermore, $ f(t, \cdot) $ is concave and there exist constants $ \chi > 0, \beta > 0 $ such that

    $ f(t,0)χ,   f(t,1)β,   0t1.
    $
    (3.8)

    Then the following results hold:

    (ⅰ) there exists $ 0 < \lambda^{*}\leq \infty $ such that the fractional boundary value problem (1.1), (1.2) has a unique positive solution $ u_{\lambda} $ in $ \tilde{K} $ for $ 0 < \lambda < \lambda^{*}, $ the fractional boundary value problem(1.1), (1.2) has no positive solution in $ \tilde{K} $ for $ \lambda\geq \lambda^{*}; $

    (ⅱ) for any $ u_{0}\in \tilde{K}, $ let

    $ u_{n}(t) = \lambda \int_{0}^{1}G(t, s)\varphi'(s)f(s, u_{n-1}(s))ds, \ n = 1, 2, 3, \cdot\cdot\cdot. $

    Then for $ 0 < \lambda < \lambda^{*}, $ we have

    $ \max\limits_{0\leq t\leq 1}|u_{n}(t)- u_{\lambda}(t)|\rightarrow 0 \ \mbox{as}\ \ n\rightarrow \infty; $

    (ⅲ) if $ 0 < \lambda < \lambda^{*}, \ \max_{0\leq t\leq 1}|u_{\lambda}(t)- u_{\lambda_{0}}(t)|\rightarrow 0 \ \mbox{as} \ \lambda\rightarrow \lambda_{0} $ and if $ 0 < \lambda_{1} < \lambda_{2} < \lambda^{*}, $ then $ u_{\lambda_{1}}(t)\leq u_{\lambda_{2}}(t), \ \ 0\leq t\leq 1 $ and $ u_{\lambda_{1}}(t)\not\equiv u_{\lambda_{2}}(t); $

    (ⅳ) $ u_{v\lambda}(t)\leq vu_{\lambda}(t) $ for $ 0 < \lambda < \lambda^{*}, \ \ 0\leq v\leq 1, \ \ 0\leq t\leq 1. $

    (ⅴ) if $ \lim_{u\rightarrow \infty}\frac{f(t, u)}{u} = 0 $ uniformly on $ [0, 1], $ then $ \lambda^{*} = \infty. $

    Proof. Define an operator $ T $ by

    $ (Tu)(t)=10G(t,s)φ(s)f(s,u(s))ds.
    $
    (3.9)

    Thus we know that $ u $ is a solution of (1.1), (1.2) if and only if $ u = \lambda Tu. $ For $ u\in K, $ in view of $ (H_{1}), (H_{2}) $ and Lemma 3.2, we can get $ Tu\in K. $ Owing to the concavity of $ f(t, \cdot), $ for $ u > 1, $ we have

    $ f\bigl(t, 1\bigr) = f\biggl(t, \frac{1}{u}u+\biggl(1-\frac{1}{u}\biggr)0\biggr)\geq \frac{1}{u}f(t, u)+\biggl(1-\frac{1}{u}\biggr)f(t, 0), $

    from (3.8), we obtain

    $ f(t, u)\leq uf(t, 1)-(u-1)f(t, 0)\leq uf(t, 1)\leq u\beta. $

    So, for $ u\in K, $ we have

    $ 0\leq \int_{0}^{1}G(t, s)\varphi'(s)f(s, u(s))ds\leq \int_{0}^{1}G(t, s)\varphi'(s)f(s, \|u\|_{E})ds\leq N_{1}g(t), \ 0\leq t\leq 1, $

    where $ N_{1} = \max_{0\leq t\leq 1}f(t, \|u\|_{E})\leq \max_{0\leq t\leq 1}f(t, \|u\|_{E}+1)\leq \beta (1+\|u\|_{E}), $ so, we have $ 0\leq (Tu)(t)\leq N_{1}g(t), $ which implies $ Tu\in X. $ Thus, $ Tu\in \tilde{K}, $ this means that $ T: \tilde{K}\rightarrow \tilde{K}. $ In view of the concavity of $ f(t, \cdot), $ we have the operator $ T $ is concave.

    From (3.9), we have $ (T\theta)(t) = \int_{0}^{1}G(t, s)\varphi'(s)f(s, 0)ds\geq \chi g(t), \ 0\leq t\leq 1. $ So, $ T\theta\in \mathring{\tilde{K}}, $ this means that $ T\theta > > \theta. $ Then all the conditions of Lemma 2.4 are satisfied. Thus from Lemma 2.4, there exists $ 0 < \lambda^{*}\leq \infty $ such that the equation $ u = \lambda Tu $ has a unique positive solution $ u_{\lambda} $ in $ \tilde{K} $ for $ 0 < \lambda < \lambda^{*}, $ the equation $ u = \lambda Tu $ has no positive solution in $ \tilde{K} $ for $ \lambda\geq \lambda^{*}.; $

    For any $ u_{0}\in \tilde{K}, $ let $ u_{n} = \lambda Tu_{n-1}, n = 1, 2, 3, \cdot\cdot\cdot, $ then we have $ u_{n}\rightarrow u_{\lambda}\ \mbox{as}\ \ n\rightarrow \infty $ for $ 0 < \lambda < \lambda^{*}, \ u_{\lambda}: (0, \lambda^{*})\rightarrow \tilde{K} $ is continuous and strongly increasing. Furthermore, $ u_{t\lambda}\leq tu_{\lambda} $ for $ 0\leq\lambda < \lambda^{*}, \ 0\leq t\leq 1;\ \lambda \rightarrow \lambda^{*}-0 $ implies $ \|u_{\lambda}\|_{X}\rightarrow \infty; $ if there exist $ v_{0}\in \tilde{K} $ and $ \lambda_{0} > 0 $ such that $ \lambda_{0}Tv_{0}\leq v_{0}, $ then $ \lambda^{*} > \lambda_{0}. $ From this results, we can get

    (ⅰ) the fractional boundary value problem (1.1), (1.2) has a unique positive solution $ u_{\lambda} $ in $ \tilde{K} $ for $ 0 < \lambda < \lambda^{*}, $ the fractional boundary value problem (1.1), (1.2) has no positive solution in $ \tilde{K} $ for $ \lambda\geq \lambda^{*}; $

    (ⅱ) for any $ u_{0}\in \tilde{K}, $ let

    $ u_{n}(t) = \lambda \int_{0}^{1}G(t, s)\varphi'(s)f(s, u_{n-1}(s))ds, \ n = 1, 2, 3, \cdot\cdot\cdot. $

    Then for $ 0 < \lambda < \lambda^{*}, $ we have

    $ \max\limits_{0\leq t\leq 1}|u_{n}(t)- u_{\lambda}(t)|\rightarrow 0 \ \mbox{as}\ \ n\rightarrow \infty; $

    (ⅲ)$ \max_{0\leq t\leq 1}|u_{\lambda}(t)- u_{\lambda_{0}}(t)|\rightarrow 0 \ \mbox{as} \ \lambda\rightarrow \lambda_{0}, $ where $ 0 < \lambda_{0} < \lambda^{*} $ and if $ 0 < \lambda_{1} < \lambda_{2} < \lambda^{*}, $ then there exists $ \delta > 0 $ such that $ u_{\lambda_{2}}(t)- u_{\lambda_{1}}(t)\geq \delta g(t), \ \ 0\leq t\leq 1, $ this means that $ u_{\lambda_{1}}(t)\leq u_{\lambda_{2}}(t), \ 0\leq t\leq 1 $ and $ u_{\lambda_{1}}(t)\not\equiv u_{\lambda_{2}}(t); $

    (ⅳ) for $ 0 < \lambda < \lambda^{*}, \ u_{\delta\lambda}(t)\leq \delta u_{\lambda}(t), \ \ 0\leq \delta\leq 1, \ \ 0\leq t\leq 1. $

    (ⅴ) Denote $ \eta = \|g\|_{E}, $ for any given $ \lambda > 0, $ by $ \lim_{u\rightarrow \infty}\frac{f(t, u)}{u} = 0, $ there exists a large $ T > 0, $ such that $ f(t, T)\leq (\eta \lambda)^{-1}T, \ \ 0\leq t\leq 1. $ Let $ v_{0}(t) = \eta^{-1}T g(t), $ from Lemma 2.5, we have

    $ λ(Tv0)(t)=λ10G(t,s)φ(s)f(s,η1Tg(s))dsλ10G(t,s)φ(s)f(s,T)ds(ηλ)1Tλg(t)=η1Tg(t)=v0(t),
    $
    $ v_{0}(t)-\lambda (Tv_{0})(t) = \int_{0}^{1}G(t, s)\varphi'(s)(\eta^{-1}T-\lambda f(s, \eta^{-1}T g(s)))ds\leq M_{2}g(t), \ \ 0\leq t\leq 1, $

    where $ M_{2} = \sup_{0\leq t\leq 1}[\eta^{-1}T+\lambda f(s, T)]\leq \eta^{-1}T+\lambda \beta (T+1), $ this means that $ v_{0}(t)-\lambda Tv_{0}\in \tilde{K}. $ That is $ \lambda Tv_{0}\leq v_{0}. $ By Lemma 2.4 (v), we have $ \lambda^{*} > \lambda, $ because of the arbitrariness of $ \lambda, $ we can get $ \lambda^{*} = \infty. $ Theorem 3.4 Assume that $ (H_{1}), (H_{2}) $ hold. Furthermore, the following conditions hold:

    (ⅰ) there exists a number $ 0 < r < 1 $ such that

    $ f(t, \omega u)\geq \omega^{r}f(t, u), \ \ 0\leq t\leq 1, \ u\geq 0, \ 0 \lt \omega \lt 1; $

    (ⅱ) there exists a number $ \xi > 0 $ such that $ f(t, 1)\leq \xi; $

    (ⅲ) $ \min_{0\leq t\leq 1}f(t, g(t)) > 0, $ where $ g(t) $ is given by (3.7).

    Then the following results hold:

    (ⅰ) for any given $ \lambda > 0, $ the fractional boundary value problem (1.1), (1.2) has a unique positive solution $ u_{\lambda} $ in $ \mathring{\tilde{K}}; $

    (ⅱ) if $ 0 < \lambda_{1} < \lambda_{2}, $ then $ u_{\lambda_{1}}(t)\leq u_{\lambda_{2}}(t), \ 0\leq t\leq 1 $ and $ u_{\lambda_{1}}(t)\not\equiv u_{\lambda_{2}}(t); $

    (ⅲ) $ \max_{0\leq t\leq 1}|u_{\lambda}(t)-u_{\lambda_{0}}(t)|\rightarrow 0 $ as $ \lambda\rightarrow \lambda_{0}; $

    (ⅳ) $ \max_{0\leq t\leq 1}|u_{\lambda}(t)|\rightarrow +\infty $ as $ \lambda\rightarrow +\infty, \ \max_{0\leq t\leq 1}|u_{\lambda}(t)|\rightarrow 0 $ as $ \lambda\rightarrow 0^{+}. $

    Proof. For $ u > 1, $ from the condition (ⅰ), we have

    $ f(t, 1) = f\biggl(t, \frac{1}{u}u\biggr)\geq \biggl(\frac{1}{u}\biggr)^{r}f(t, u), $

    thus, we get

    $ f(t, u)\leq u^{r}f(t, 1)\leq \xi u^{r}. $

    So, for $ u\in K, $ we have

    $ 0\leq \int_{0}^{1}G(t, s)\varphi'(s)f(s, u(s))ds\leq \int_{0}^{1}G(t, s)\varphi'(s)f(s, \|u\|_{E})ds\leq M_{3}g(t), \ 0\leq t\leq 1, $

    where $ M_{3} = \max_{0\leq t\leq 1}f(t, \|u\|_{E})\leq \max_{0\leq t\leq 1}f(t, \|u\|_{E}+1)\leq \xi (1+\|u\|_{E})^{r}, $

    Thus, $ 0\leq Tu(t)\leq M_{3}g(t), \ \ 0\leq t\leq 1, $ this means that $ T: \tilde{K}\rightarrow \tilde{K}. $

    For $ u\in \mathring{\tilde{K}}, $ there exists $ \delta > 0 $ such that $ u(t)\geq \delta g(t)\geq 0, \ \ 0\leq t\leq 1. $ Let $ \omega\in (0, 1) $ such that $ \omega < \delta, $ we have

    $ (Tu)(t)=10G(t,s)φ(s)f(s,u(s))ds10G(t,s)φ(s)f(s,δg(s))ds10G(t,s)φ(s)f(s,ωg(s))dsωr10G(t,s)φ(s)f(s,g(s))ds.
    $

    Set $ a = \min_{0\leq t\leq 1}f(t, g(t)), $ then $ a > 0 $ and $ (Tu)(t)\geq a\omega^{r}g(t), \ 0\leq t\leq 1, $ so $ T: \mathring{\tilde{K}}\rightarrow \mathring{\tilde{K}}. $ From $ f(t, \cdot) $ is increasing, we can get that $ T $ is increasing. Moreover, for $ u\in \mathring{\tilde{K}} $ and $ \omega\in (0, 1), $ we have

    $ T(\omega u)(t) = \int_{0}^{1}G(t, s)\varphi'(s)f(s, \omega u(s))ds\geq \int_{0}^{1}G(t, s)\varphi'(s)\omega^{r}f(s, u(s))ds = \omega^{r}(Tu)(t). $

    In the following, we consider the operator equation

    $ (Tu)(t)=ρu(t).
    $
    (3.10)

    Then all the conditions of Lemma 2.6 are satisfied. Thus, from Lemma 2.6, for any given $ \rho > 0, $ (3.10) has a unique solution $ u_{\rho} $ in $ \mathring{\tilde{K}}, $ $ u_{\rho} $ is strongly decreasing, $ u_{\rho} $ is continuous, this means that $ \|u_{\rho}-u_{\rho_{0}}\|\rightarrow 0 $ as $ \rho\rightarrow \rho_{0} $ and furthermore, $ \lim_{\rho\rightarrow +\infty}\|u_{\rho}\| = 0, \ \ \ \lim_{\rho\rightarrow 0^{+}}\|u_{\rho}\| = +\infty. $ Set $ \lambda = \frac{1}{\rho}, \ \lambda_{0} = \frac{1}{\rho_{0}}, \ \lambda_{1} = \frac{1}{\rho_{1}}, \ \lambda_{2} = \frac{1}{\rho_{2}}. $ Then (3.10) is equivalent to $ u(t) = \lambda (Tu)(t). $ As a result, we can get

    (ⅰ) for any given $ \lambda > 0, $ the fractional boundary value problem (1.1), (1.2) has a unique positive solution $ u_{\lambda} $ in $ \mathring{\tilde{K}}; $

    (ⅱ) if $ 0 < \lambda_{1} < \lambda_{2}, $ there exists $ \delta > 0 $ such that $ u_{\lambda_{2}}(t)-u_{\lambda_{1}}(t)\geq \delta g(t) $ which implies $ u_{\lambda_{1}}(t)\leq u_{\lambda_{2}}(t) $ and $ u_{\lambda_{1}}(t)\not\equiv u_{\lambda_{2}}(t), \ 0\leq t\leq 1; $

    (ⅲ) $ \max_{0\leq t\leq 1}|u_{\lambda}(t)-u_{\lambda_{0}}(t)|\rightarrow 0 $ as $ \lambda\rightarrow \lambda_{0}; $

    (ⅳ) $ \max_{0\leq t\leq 1}|u_{\lambda}(t)|\rightarrow +\infty $ as $ \lambda\rightarrow +\infty, \ \ \max_{0\leq t\leq 1}|u_{\lambda}(t)|\rightarrow 0 $ as $ \lambda\rightarrow 0^{+}. $

    Example 4.1 Consider the fractional boundary value problem

    $ D32, t20+u(t)+λ(t2+1+u13)=0,   0<t<1,
    $
    (4.1)
    $ u(0)=0,   u(1)=12u(12).
    $
    (4.2)

    We note that $ \alpha = \frac{3}{2}, \beta = \frac{1}{2}, \eta = \frac{1}{2}, \varphi(t) = \frac{t}{2}, f(t, u) = t^{2}+1+u^{\frac{1}{3}}. $ Thus, we have $ f(t, u):[0, 1]\times R^{+}\rightarrow R^{+} $ is continuous and increasing in $ u, f(t, u) $ is concave. We can check that conditions $ (H_{1}), (H_{2}) $ are satisfied. Choose $ \alpha = 1, \beta = 3 $ and $ f(t, u) $ satisfy

    $ f(t, 0) = t^{2}+1\geq 1, \ \ f(t, 1) = t^{2}+2\leq 3, \ \ 0\leq t\leq 1, $
    $ \lim\limits_{u\rightarrow \infty}\frac{f(t, u)}{u} = \lim\limits_{u\rightarrow \infty}\frac{t^{2}+1+u^{\frac{1}{3}}}{u} = 0. $

    Then all the assumptions of Theorem 3.3 are satisfied. Thus from Theorem 3.3, we have $ \lambda^{*} = \infty, $ and the following conclusions hold:

    (ⅰ) for $ \lambda > 0, $ the fractional boundary value problem (4.1), (4.2) has a unique positive solution $ u_{\lambda} $ in $ \tilde{K}; $

    (ⅱ) for any $ u_{0}\in \tilde{K}, $ let

    $ u_{n}(t) = \lambda \int_{0}^{1}G(t, s)\varphi'(s)f(s, u_{n-1}(s))ds, \ n = 1, 2, 3, \cdot\cdot\cdot. $

    Then for $ \lambda > 0, $ we have

    $ \max\limits_{0\leq t\leq 1}|u_{n}(t)- u_{\lambda}(t)|\rightarrow 0 \ \mbox{as}\ \ n\rightarrow \infty; $

    (ⅲ) $ \max_{0\leq t\leq 1}|u_{\lambda}(t)- u_{\lambda_{0}}(t)|\rightarrow 0 \ \mbox{as} \ \lambda\rightarrow \lambda_{0} $ and if $ 0 < \lambda_{1} < \lambda_{2} < +\infty, $ then $ u_{\lambda_{1}}(t)\leq u_{\lambda_{2}}(t), \ \ 0\leq t\leq 1 $ and $ u_{\lambda_{1}}(t)\not\equiv u_{\lambda_{2}}(t); $

    (ⅳ) $ u_{v\lambda}(t)\leq vu_{\lambda}(t) $ for $ 0 < \lambda < +\infty, \ \ 0\leq v\leq 1, \ \ 0\leq t\leq 1. $

    In this paper, we discussed a class of fractional derivative with respect to another function called $ \varphi $-Riemann-Liouville fractional derivative. The new derivative herein generalizes the classical definitions of derivatives by choosing particular forms of $ \varphi(t). $ The main advantage of our paper is we not only develop the existence and uniqueness of positive solutions to the fractional boundary value problem, but also give some properties of positive solutions to the problem.

    We express our sincere thanks to the anonymous reviewers for their valuable comments and suggestions. This work is supported by the Natural Science Foundation of Tianjin (No.(19JCYBJC30700)).

    The authors declare no conflict of interest in this paper.

    [1] O'Regan B, Gratzel M (1991) A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 353: 734–740.
    [2] Mathew S, Yella A, Gao P, et al. (2014) Dye-sensitized solar cells with 13% efficiency achieved through the molecular engineering of porphyrin sensitizers. Nat Chem 6: 242–247. doi: 10.1038/nchem.1861
    [3] Kojima A, Teshima K, Shirai Y, et al. (2009) Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J Am Chem Soc 131: 6050–6051. doi: 10.1021/ja809598r
    [4] Lee MM, Teuscher J, Miyasaka T, et al. (2012) Efficient hybrid solar cells based on meso-superstructured organometal halide perovskites. Science 338: 643–647. doi: 10.1126/science.1228604
    [5] Burschka J, Pellet N, Moon S-J, et al. (2013) Sequential deposition as a route to high-performance perovskite-sensitized solar cells. Nature 499: 316–320. doi: 10.1038/nature12340
    [6] Yang WS, Noh JH, Jeon NJ, et al. (2015) High-performance photovoltaic perovskite layers fabricated through intramolecular exchange. Science 348: 1234–1237. doi: 10.1126/science.aaa9272
    [7] Xu F, Sun L (2011) Solution-derived ZnO nanostructures for photoanodes of dye-sensitized solar cells. Energy Environ Sci 4: 818–841. doi: 10.1039/C0EE00448K
    [8] Zhang Q, Cao G (2011) Nanostructured photoelectrodes for dye-sensitized solar cells. Nano Today 6: 91–109. doi: 10.1016/j.nantod.2010.12.007
    [9] Strehlow WH, Cook EL (1973) Compilation of energy band gaps in elemental and binary compound semiconductors and insulators. J Phys Chem Ref Data 2: 163–200. doi: 10.1063/1.3253115
    [10] Scanlon DO, Dunnill CW, Buckeridge J, et al. (2013) Band alignment of rutile and anatase TiO2. Nat Mater 12: 798–801. doi: 10.1038/nmat3697
    [11] Chou TP, Zhang Q, Fryxell GE, et al. (2007) Hierarchically structured ZnO film for dye-sensitized solar cells with enhanced energy conversion efficiency. Adv Mater 19: 2588–2592. doi: 10.1002/adma.200602927
    [12] Dittrich T, Lebedev EA, Weidmann J (1998) Electron drift mobility in porous TiO2 (anatase). Phys Status Solidi A 165: R5–R6.
    [13] Redmond G, Fitzmaurice D, Graetzel M (1994) Visible light sensitization by cis-bis(thiocyanato)bis(2,2'-bipyridyl-4,4'-dicarboxylato)ruthenium(II) of a transparent nanocrystalline ZnO film prepared by sol-gel techniques. Chem Mater 6: 686–691.
    [14] Hoyer P, Weller H (1995) Potential-dependent electron injection in nanoporous colloidal ZnO films. J Phys Chem 99: 14096–14100. doi: 10.1021/j100038a048
    [15] Rensmo H, Keis K, Lindström H, et al. (1997) High light-to-energy conversion efficiencies for solar cells based on nanostructured ZnO electrodes. J Phys Chem B 101: 2598–2601. doi: 10.1021/jp962918b
    [16] van de Lagemaat J, Frank AJ (2001) Nonthermalized electron transport in dye-sensitized nanocrystalline TiO2 films:  Transient photocurrent and random-walk modeling studies. J Phys Chem B 105: 11194–11205. doi: 10.1021/jp0118468
    [17] Frank A, Kopidakis N, Lagemaat Jvd (2004) Electrons in nanostructured TiO2 solar cells: transport, recombination and photovoltaic properties. Coord Chem Rev 248: 1165–1179. doi: 10.1016/j.ccr.2004.03.015
    [18] Willis RL, Olson C, O'Regan B, et al. (2002) Electron dynamics in nanocrystalline ZnO and TiO2 films probed by potential step chronoamperometry and transient absorption spectroscopy. J Phys Chem B 106: 7605–7613. doi: 10.1021/jp020231n
    [19] Quintana M, Edvinsson T, Hagfeldt A, et al. (2007) Comparison of dye-sensitized ZnO and TiO2 solar cells:  Studies of charge transport and carrier lifetime. J Phys Chem C 111: 1035–1041. doi: 10.1021/jp065948f
    [20] Kopidakis N, Benkstein KD, van de Lagemaat J, et al. (2003) Transport-limited recombination of photocarriers in dye-sensitized nanocrystalline TiO2 solar cells. J Phys Chem B 107: 11307–11315.
    [21] Martinson ABF, McGarrah JE, Parpia MOK, et al. (2006) Dynamics of charge transport and recombination in ZnO nanorod array dye-sensitized solar cells. Phys Chem Chem Phys 8: 4655–4659. doi: 10.1039/b610566a
    [22] Baxter JB, Walker AM, Ommering Kv, et al. (2006) Synthesis and characterization of ZnO nanowires and their integration into dye-sensitized solar cells. Nanotechnol 17: S304. doi: 10.1088/0957-4484/17/11/S13
    [23] Galoppini E, Rochford J, Chen H, et al. (2006), Fast electron transport in metal organic vapor deposition grown dye-sensitized ZnO nanorod solar cells. J Phys Chem B 110: 16159–16161.
    [24] Law M, Greene LE, Johnson JC, et al. (2005) Nanowire dye-sensitized solar cells. Nat Mater 4: 455–459. doi: 10.1038/nmat1387
    [25] Noack V, Weller H, Eychmüller A (2002) Electron transport in particulate ZnO electrodes:  A simple approach. J Phys Chem B 106: 8514–8523. doi: 10.1021/jp0200270
    [26] Qiu J, Li X, Zhuge F, et al. (2010) Solution-derived 40 µm vertically aligned ZnO nanowire arrays as photoelectrodes in dye-sensitized solar cells. Nanotechnol 21: 195602. doi: 10.1088/0957-4484/21/19/195602
    [27] Ku C-H, Wu J-J (2007) Chemical bath deposition of ZnO nanowire–nanoparticle composite electrodes for use in dye-sensitized solar cells. Nanotechnol 18: 505706. doi: 10.1088/0957-4484/18/50/505706
    [28] Xu F, Dai M, Lu Y, et al. (2010) Hierarchical ZnO nanowire−nanosheet architectures for high power conversion efficiency in dye-sensitized solar cells. J Phys Chem C 114: 2776–2782. doi: 10.1021/jp910363w
    [29] Xu C, Wu J, Desai UV, et al. (2011) Multilayer assembly of nanowire arrays for dye-sensitized solar cells. J Am Chem Soc 133: 8122–8125. doi: 10.1021/ja202135n
    [30] Seow ZLS, Wong ASW, Thavasi V, et al. (2009) Controlled synthesis and application of ZnO nanoparticles, nanorods and nanospheres in dye-sensitized solar cells. Nanotechnol 20: 045604. doi: 10.1088/0957-4484/20/4/045604
    [31] Cakir AC, Erten-Ela S (2012) Comparison between synthesis techniques to obtain ZnO nanorods and its effect on dye sensitized solar cells. Adv Powder Technol 23: 655–660. doi: 10.1016/j.apt.2011.08.003
    [32] Hosni M, Kusumawati Y, Farhat S, et al. (2014), Effects of oxide nanoparticle size and shape on electronic structure, charge transport, and recombination in dye-sensitized solar cell photoelectrodes. J Phys Chem C 118: 16791–16798.
    [33] Cao HL, Qian XF, Gong Q, et al. (2006) Shape- and size-controlled synthesis of nanometre ZnO from a simple solution route at room temperature. Nanotechnol 17: 3632. doi: 10.1088/0957-4484/17/15/002
    [34] Bai S, Hu J, Li D, et al. (2011) Quantum-sized ZnO nanoparticles: Synthesis, characterization and sensing properties for NO2. J Mater Chem 21: 12288–12294. doi: 10.1039/c1jm11302j
    [35] Hu X, Masuda Y, Ohji T, et al. (2008) Micropatterning of ZnO nanoarrays by forced hydrolysis of anhydrous zinc acetate. Langmuir 24: 7614–7617. doi: 10.1021/la8006348
    [36] Tan ST, Sun XW, Zhang XH, et al. (2006) Cluster coarsening in zinc oxide thin films by postgrowth annealing. J Appl Phys 100: 033502. doi: 10.1063/1.2218468
    [37] Hill JJ, Haller K, Gelfand B, et al. (2010) Eliminating capillary coalescence of nanowire arrays with applied electric fields. ACS Appl Mater Interfaces 2: 1992–1998. doi: 10.1021/am100290z
    [38] Sutton LE, HJM Bowen (1958) Tables of Interatomic Distances and Configuration in Molecules and Ions, London: The Chemical Society.
    [39] Spiro M, Creeth AM (1990) Tracer diffusion coefficients of I-, I3-, Fe2+ and Fe3+ at low temperatures. J Chem Soc Faraday Trans 86: 3573–3576. doi: 10.1039/ft9908603573
    [40] Lin C, Tsai FY, Lee FH, et al. (2009) Enhanced performance of dye-sensitized solar cells by an Al2O3 charge-recombination barrier formed by low-temperature atomic layer deposition. J Mater Chem 19: 2999–3003. doi: 10.1039/b819337a
    [41] Mozer AJ, Wagner P, Officer DL, et al. (2008) The origin of open circuit voltage of porphyrin-sensitised TiO2 solar cells. Chem Commun 4741–4743.
    [42] Ko SB, Cho AN, Kim MJ, et al., (2012) Alkyloxy substituted organic dyes for high voltage dye-sensitized solar cell: Effect of alkyloxy chain length on open-circuit voltage. Dyes Pigments 94: 88–98. doi: 10.1016/j.dyepig.2011.10.014
    [43] Kalyanasundaram K (2009) Dye-sensitized solar cells. Boca Raton, Fla: CRC.
  • This article has been cited by:

    1. Mondher Benjemaa, Fatma Jerbi, On differential equations involving the ψψ
    ‐shifted fractional operators, 2023, 46, 0170-4214, 3371, 10.1002/mma.8697
    2. Golnaz Pakgalb, Mohammad Jahangiri Rad, Ali Salimi Shamloo, Majid Derafshpour, Existence and uniqueness of a positive solutions for the product of operators, 2022, 7, 2473-6988, 18853, 10.3934/math.20221038
    3. Ricardo Almeida, Euler–Lagrange-Type Equations for Functionals Involving Fractional Operators and Antiderivatives, 2023, 11, 2227-7390, 3208, 10.3390/math11143208
  • Reader Comments
  • © 2016 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(7223) PDF downloads(1612) Cited by(5)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog