13
Dynamics of silica cell wall morphogenesis in the diatom Cyclotella cryptica: Substructure formation and the role of microfilaments Benoit Tesson, Mark Hildebrand * Marine Biology Research Division, Scripps Institution of Oceanography, University of California San Diego, 9500 Gilman Dr., La Jolla, CA 92093-0202, USA article info Article history: Received 15 June 2009 Received in revised form 22 August 2009 Accepted 27 August 2009 Available online 1 September 2009 Keywords: Diatom Biomineralization Nanostructure Materials Microscopy abstract Diatoms are unicellular algae that make cell walls out of silica with highly ornate features on the nano- to microscale. The complexity and variety of diatom cell wall structures exceeds those possible with syn- thetic materials chemistry approaches. Understanding the design and assembly processes involved in diatom silicification should provide insight into patterning on the unicellular level, and information for biomimetic approaches for materials synthesis. In this report we examine the formation of distinct cell wall structures (valves and girdle bands) in the diatom Cyclotella cryptica by high resolution imaging using SEM, AFM, and fluorescence microscopy. Special attention was paid to imaging structural interme- diates, which provided insight into the underlying design and assembly principles involved. Distinct stages in valve formation were identified, indicating a transition from a fractally organized structure to a dynamic pathway-dependent process. Substructures in the valves appeared to be pre-positioned prior to complete silicification, suggesting that organics responsible for these structures were pre-assembled and put in place. Microtubules and microfilamentous actin play significant roles in the positioning pro- cess, and actin is also important in the pathway-dependent expansion of the front of silicification. Our results indicate that even though all silica structures in C. cryptica are made of assemblies of nanopartic- ulate silica, control of meso- and microscale structure occurs on a higher order. It is apparent that dia- toms integrate bottom up and top down control and synthesis mechanisms to form the diversity of structures possible. Ó 2009 Elsevier Inc. All rights reserved. 1. Introduction Diatoms are perhaps the most impressive group of silicifying organisms, with more than 10,000 species and an even greater number of different structures in their cell walls formed out of sil- ica on the nano- to microscale. In addition to their silica structures, diatom genomic sequencing (Armbrust et al., 2004; Bowler et al., 2008) and genetic manipulation (transgenic) capabilities (Apt et al., 1996; Dunahay et al., 1995; Hildebrand, 2008; Kröger and Poulsen, 2008) have rendered diatoms as model organisms for the study of nanoscale pattern formation and silica mineralization at the single cell level. The cell biology of diatom silicification has a stand-alone interest; however, because diatoms are capable of making structures that exceed those possible using current mate- rials synthesis approaches, they are being looked at as a possible biologically-based system to generate nanomaterials for techno- logical applications (Hildebrand, 2008; Losic et al., 2009; Sandhage et al., 2005). This could entail either the direct use of materials de- rived from diatom silica (Sandhage et al., 2005), or the application of lessons learned from the study of diatom silicification to biomi- metic approaches (Losic et al., 2009). In either case, it will be essen- tial to understand at the molecular level the processes by which diatoms control the formation of silica from the microscale to the nanoscale. A primary concept in biomineralization is the need to generate a defined space in order to control the growth and shape of the structure being produced. This is particularly true with amorphous minerals such as silica. By definition, amorphous minerals do not possess structural order at the molecular level, which implies that there is no preferential growth direction (Addadi and Weiner, 1992). In the case of silica, its polymerization can be induced rela- tively easily (Iler, 1979), but forming a specific shape requires con- trol at a different level. Some basic concepts and components involved in diatom cell wall formation have been described. There are two major struc- tural features comprising the diatom cell wall (also called the frus- tule), which has the general structure of a petri dish. The valves are the distinctive structures characteristic of a given diatom species and form the extreme upper and lower portions of the wall, and the girdle bands are most commonly thin strips of silica which encircle the cell, and constitute the side walls and provide overlap 1047-8477/$ - see front matter Ó 2009 Elsevier Inc. All rights reserved. doi:10.1016/j.jsb.2009.08.013 * Corresponding author. E-mail address: [email protected] (M. Hildebrand). Journal of Structural Biology 169 (2010) 62–74 Contents lists available at ScienceDirect Journal of Structural Biology journal homepage: www.elsevier.com/locate/yjsbi

Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

Journal of Structural Biology 169 (2010) 62–74

Contents lists available at ScienceDirect

Journal of Structural Biology

journal homepage: www.elsevier .com/ locate/yjsbi

Dynamics of silica cell wall morphogenesis in the diatom Cyclotella cryptica:Substructure formation and the role of microfilaments

Benoit Tesson, Mark Hildebrand *

Marine Biology Research Division, Scripps Institution of Oceanography, University of California San Diego, 9500 Gilman Dr., La Jolla, CA 92093-0202, USA

a r t i c l e i n f o a b s t r a c t

Article history:Received 15 June 2009Received in revised form 22 August 2009Accepted 27 August 2009Available online 1 September 2009

Keywords:DiatomBiomineralizationNanostructureMaterialsMicroscopy

1047-8477/$ - see front matter � 2009 Elsevier Inc. Adoi:10.1016/j.jsb.2009.08.013

* Corresponding author.E-mail address: [email protected] (M. Hildeb

Diatoms are unicellular algae that make cell walls out of silica with highly ornate features on the nano- tomicroscale. The complexity and variety of diatom cell wall structures exceeds those possible with syn-thetic materials chemistry approaches. Understanding the design and assembly processes involved indiatom silicification should provide insight into patterning on the unicellular level, and information forbiomimetic approaches for materials synthesis. In this report we examine the formation of distinct cellwall structures (valves and girdle bands) in the diatom Cyclotella cryptica by high resolution imagingusing SEM, AFM, and fluorescence microscopy. Special attention was paid to imaging structural interme-diates, which provided insight into the underlying design and assembly principles involved. Distinctstages in valve formation were identified, indicating a transition from a fractally organized structure toa dynamic pathway-dependent process. Substructures in the valves appeared to be pre-positioned priorto complete silicification, suggesting that organics responsible for these structures were pre-assembledand put in place. Microtubules and microfilamentous actin play significant roles in the positioning pro-cess, and actin is also important in the pathway-dependent expansion of the front of silicification. Ourresults indicate that even though all silica structures in C. cryptica are made of assemblies of nanopartic-ulate silica, control of meso- and microscale structure occurs on a higher order. It is apparent that dia-toms integrate bottom up and top down control and synthesis mechanisms to form the diversity ofstructures possible.

� 2009 Elsevier Inc. All rights reserved.

1. Introduction

Diatoms are perhaps the most impressive group of silicifyingorganisms, with more than 10,000 species and an even greaternumber of different structures in their cell walls formed out of sil-ica on the nano- to microscale. In addition to their silica structures,diatom genomic sequencing (Armbrust et al., 2004; Bowler et al.,2008) and genetic manipulation (transgenic) capabilities (Aptet al., 1996; Dunahay et al., 1995; Hildebrand, 2008; Kröger andPoulsen, 2008) have rendered diatoms as model organisms forthe study of nanoscale pattern formation and silica mineralizationat the single cell level. The cell biology of diatom silicification has astand-alone interest; however, because diatoms are capable ofmaking structures that exceed those possible using current mate-rials synthesis approaches, they are being looked at as a possiblebiologically-based system to generate nanomaterials for techno-logical applications (Hildebrand, 2008; Losic et al., 2009; Sandhageet al., 2005). This could entail either the direct use of materials de-rived from diatom silica (Sandhage et al., 2005), or the application

ll rights reserved.

rand).

of lessons learned from the study of diatom silicification to biomi-metic approaches (Losic et al., 2009). In either case, it will be essen-tial to understand at the molecular level the processes by whichdiatoms control the formation of silica from the microscale to thenanoscale.

A primary concept in biomineralization is the need to generatea defined space in order to control the growth and shape of thestructure being produced. This is particularly true with amorphousminerals such as silica. By definition, amorphous minerals do notpossess structural order at the molecular level, which implies thatthere is no preferential growth direction (Addadi and Weiner,1992). In the case of silica, its polymerization can be induced rela-tively easily (Iler, 1979), but forming a specific shape requires con-trol at a different level.

Some basic concepts and components involved in diatom cellwall formation have been described. There are two major struc-tural features comprising the diatom cell wall (also called the frus-tule), which has the general structure of a petri dish. The valves arethe distinctive structures characteristic of a given diatom speciesand form the extreme upper and lower portions of the wall, andthe girdle bands are most commonly thin strips of silica whichencircle the cell, and constitute the side walls and provide overlap

Page 2: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

B. Tesson, M. Hildebrand / Journal of Structural Biology 169 (2010) 62–74 63

between the two halves of the cell. Observational studies haveidentified three scales of structure formation in diatoms (Davisand Hildebrand, 2007; Hildebrand et al., 2006; Pickett-Heapset al., 1990); (1) the nanoscale, which involves the initial polymer-ization of silica from soluble precursors and which manifests as sil-ica structures with nanometer-sized features, (2) the microscale,which relates to the overall shape of the silica structure formedwithin the confines of the silica deposition vesicle (SDV), whichis a membrane-bound intracellular compartment wherein silicifi-cation occurs (Drum and Pankratz, 1964; Pickett-Heaps et al.,1990), and (3) the mesoscale, in which nanoscale polymerizationproducts are organized at a higher level within the confines ofthe SDV to form intermediate-sized silica structures. Mesoscalestructures are commonly the features most distinct betweenspecies.

Molecular components involved in polymerization and struc-ture formation have been described for the nano- and microscales.At the nanoscale, these include long chain polyamines (LCPAs),which catalyze silica polymerization (Kröger et al., 2000), silaffins,which are highly post-translationally modified (poly)peptides thatplay either a catalytic or regulatory role in silica polymerization byorganizing LCPAs via electrostatic interactions (Kröger et al., 1999,2001, 2002), and silacidins, which are polyanionic polypeptidesthat are also involved in the organization of silica polymerizationdeterminants (Wenzl et al., 2008). Nanoscale silica morphology isaffected by the type of silaffin and ratio of silaffin to polyamineto silacidin, however well-organized higher order structure doesnot spontaneously form from these molecules (Kröger et al.,2000, 2002; Wenzl et al., 2008). On the microscale, the SDV mem-brane (the silicalemma) plays a critical role in silica structuring,having been shown to be shaped and molded both actively andpassively to form structure (Pickett-Heaps et al., 1990). To dateno organic components definitively shown to be part of the silica-lemma have been characterized. A close association has beenshown between the SDV and the cytoskeletal elements actin andmicrotubules, and these protein assemblies play a critical role inshaping microscale structure (Pickett-Heaps et al., 1990; van deMeene and Pickett-Heaps, 2002, 2004). Inhibitors of cytoskeletonformation can generate substantial changes in diatom silica struc-ture (Cohn et al., 1989; Schmid, 1984). Direct visualization hasshown that microtubules are generally found along rib structuresin diatom silica, whereas actin is associated with the active frontof silicification (Pickett-Heaps et al., 1990; van de Meene and Pick-ett-Heaps, 2002, 2004). Fluorescence microscopy to localize actinand microtubules has only been done on two diatom species, andin both cases, microtubules were seen aligned in the direction ofgrowth of the SDV just ahead of the silicification front, and actinformed a ring structure that defined the diameter of the SDV,and expanded in diameter with increasing expansion of the SDV(van de Meene and Pickett-Heaps, 2002, 2004).

From what has been described, it is clear that diatom silicifica-tion involves both ‘‘bottom up” (generation of larger-scale struc-ture from assembly of smaller precursors), and ‘‘top down”(shaping of smaller scale structure from a larger scale) processesto form the myriad of structures possible. The scale at which thebottom up and top down processes interact is the mesoscale,which is the least well characterized scale of diatom silicification.There are several models for how mesoscale structure may form.It has been proposed that (1) organic molecules serve as organizersof polymerization determinants inside the SDV (Hildebrand, 2008;Hildebrand et al., 2009b; Robinson and Sullivan, 1987), (2) the pro-cess occurs through diffusion limited aggregation where the role oforganics is limited to providing a confined space (Parkinson et al.,1999), or (3) self assembly of silaffins and polyamines and silaci-dins (Sumper and Kröger, 2004; Wenzl et al., 2008) or phase sepa-ration processes involving polyamine droplets (Lenoci and Camp,

2008; Sumper, 2002) generate higher order structure. Observa-tional data has resulted in the generation of models for mesoscaleformation in which cytoskeletal elements located outside the SDVinterface with membrane proteins spanning the silicalemma thatposition silica polymerization determinants inside (Pickett-Heapset al., 1990; Robinson and Sullivan, 1987). These models not onlydescribe a possible mechanism for mesoscale structure formation,but explain why the forming valve is anchored to one side of theSDV and only expands on the other side, which is commonly seen(Hildebrand et al., 2006; Schmid and Schulz, 1979; Schmid andVolcani, 1983). It is important to consider that, given the complex-ity of diatom silica structure, multiple mechanisms may be in-volved in formation of mesoscale structure, indeed in a recentreport, both internal templating and confinement models couldbe used to explain formation of two different structures in Thalass-iosira pseudonana (Hildebrand et al., 2009a).

Regardless of the model proposed, the predicted process ofstructure formation must correspond with observational data.Thus, there is continuing value for detailed microscopic analysisof diatom silica structure formation, especially with the applicationof advanced high resolution imaging techniques. Since silicificationis a non-equilibrium process, knowledge of the dynamics of thephenomenon and characterization of intermediate stages duringwall formation is essential for placing the role of different compo-nents in the correct context.

Recently, a detailed examination of different stages in formationof the valve of T. pseudonana was presented (Hildebrand et al.,2006). After formation of a ‘‘base layer” consisting of depositionof radially-oriented ribs of silica to form the outline of the valve,additional deposition occurred only on one side of the ribs (the dis-tal side) resulting in rigidification of the structure (Hildebrandet al., 2006). The T. pseudonana valve has a relatively simple struc-ture, and yet characterization of its formation indicated a carefullycontrolled assembly process. Examination of structure formationprocesses in species with valves having increasingly complicatedshapes should be beneficial for understanding the underlying gen-eral design and assembly principles that diatoms utilize. Cyclotellacryptica is an attractive next species for this purpose because of itsdistinctive valve structure (described herein), it’s relatedness to T.pseudonana in the same class of the Thalassiosirales, and the factthat it is genetically manipulable (Dunahay et al., 1995), whichshould enable application of transgenic techniques to study silicastructure formation. In this work we describe in detail the dynam-ics of formation of different components of the frustule of C. cryp-tica using SEM, AFM, and fluorescence microscopy. Our resultshighlight the apparent pre-positioning of organic componentsprior to complete silicification, and the role of the cytoskeletonin both positioning and in the dynamic growth of silica structure.

2. Materials and methods

2.1. Diatom culture conditions

Cyclotella cryptica strain CCMP332 was cultured in NEPC med-ium (http://www.botany.ubc.ca/cccm/NEPCC/esaw.html) at 18 �Cunder continuous light. Cultures were synchronized using a siliconstarvation/replenishment procedure developed for T. pseudonana(Hildebrand et al., 2007), and although the cell cycle arrest pointresulting from silicon starvation differed from T. pseudonana, cellsundergoing valve synthesis were still enriched. When used, cyto-chalasin D (200 lM) and colchicine (12.5 mM) were prepared inDMSO and added to the medium at 3 and 50 lM, respectively.Oryzalin (2.9 mM) was dissolved in water and added to the med-ium at 0.2 lM.

Page 3: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

64 B. Tesson, M. Hildebrand / Journal of Structural Biology 169 (2010) 62–74

2.2. Sample preparation for SEM and AFM

Diatom frustules were cleaned by acid treatment. Ten millilitersof diatom culture was harvested by centrifugation at 4000 rpm for4 min, rinsed once with 2.3% NaCl and frozen at �20 �C. Acid treat-ment of cell walls was done by boiling cells in 1 ml concentratedsulfuric acid for 10 min, cooling, and then adding 20 mg KNO3, thenboiling an additional 10 min. Frustules were then washed usingcentrifugation three times with ultra pure water. For observationof frustules in cross-section, cleaned frustules were sonicatedtwo times for 30 s each in pulsed mode (Vibra cell, Sonics andMaterials, Newtown, CT, USA).

For SEM examination, samples were coated with gold/palla-dium and observed with an FEI Quanta 600 (FEI Company, Hills-boro, OR, USA) scanning electron microscope at the ScrippsInstitution of Oceanography Unified Laboratory Facility. High reso-lution imaging was performed on a Philips XL 30 ESEM (UCSD, Ca-lit2 Nano3 facilities).

For AFM imaging, frustules were mounted on poly-L-lysinecoated slides and imaged in air. Images were acquired in AC modeusing a silicon cantilever with a spring constant of 42 N/m(AC160TS, Olympus) on an Asylum MFP-3D-BIO Atomic ForceMicroscope coupled with an Olympus inverted fluorescent micro-scope. AFM images were processed using WSxM 4.0 software (Hor-cas et al., 2007).

Hydrofluoric acid etching was performed on acid cleaned frus-tules by suspending in 0.05 M HF for 5–10 min and then washingby centrifugation several times with ultrapure water.

2.3. Sample preparation for fluorescence microscopy

Silica incorporation was visualized by addition in the culturemedium of 100 ng ml�1 of PDMPO ([2-(4-pyridyl)-5-((4-(2-dim-ethylaminoethylamino-carbamoyl)methoxy)phenyl)oxazole](Invitrogen Corp., La Jolla, CA) prior to silicate replenishment (Shi-mizu et al., 2001). Actin staining followed the procedure of van deMeene and Pickett-Heaps (2002), where cells were fixed in 2%formaldehyde prepared in Actin Stabilizing Buffer (ASB – 10 mMPIPES/10 mM EGTA/5 mM MgSO4/pH 6.9) and containing 100 lMMBS (m-maleimidobenzoyl N-hydroxysuccinimide ester) and 1%NaCl at 4 �C, rinsed twice in ASB buffer and stained with rhoda-mine phalloidin (Invitrogen Corp., La Jolla, CA) diluted at 1/100 inASB buffer overnight at 4 �C. Samples were then rinsed twice inASB prior to observation using a Zeiss AxioObserver invertedmicroscope equipped with an Apotome (Carl Zeiss Microimaging,Inc., Thornwood, NY, USA). The filter set used for PDMPO was Zeiss#21HE (Ex 387/15 nm, FT 409, Em 510/90 nm) and for rhodaminephalloidin, Zeiss #43HE (Ex 550/25 nm, FT 570 nm, Em 605/70 nm), respectively. With these filters and under the exposuresused, no chlorophyll autofluorescence was visible. Chlorophyllwas imaged using Zeiss filter set #16 (Ex 485/20 nm, FT 510 nm,Em 515 nm LP). Images were acquired with 63�/1.4 objective oilimmersion plan APO and treated using Axiovision 4.7.2 software.Presented images are from 3D reconstructions (except when spec-ified), and presented in girdle band view.

3. Results

3.1. Characteristics of the valve and its formation

The valve of C. cryptica is circular and averages 7–8 lm in diam-eter. The distal valve surface (away from the cell interior) consistsof a radially-oriented pattern of alternating wide ribs and grid-likearrangements of pores (Fig. 1a and b). The central portion of thevalve (Fig. 1c), constituting about half of the valve diameter, has

low relief features and is relatively flat compared with the grid-likeportion at the periphery of the valve (Fig. 1b). The overall reliefshape of the valve in side view is undulated near the edges, withpore fields being rounded up above the flat ribs and central region(Fig. 1d). Fultoportulae are the moderately extended tubular struc-tures located on the valve rim at the ends of the wide ribs and inbetween the pore fields (Fig. 1d, arrows). Most, but not all, of thewide ribs contain fultoportulae (Fig. 1e and f). Located in the cen-tral portion of the valve are one or two fultoportula which have aninternal structure similar to the fultoportulae on the rim of thevalve (Fig. 1e), but protrude less from the distal valve surface(Fig. 1a and c, arrows). The central fultoportula is more easily vis-ible in the proximal surface valve view (Fig. 1e, arrow). The view ofthe proximal surface highlights the predominance of the wide ribstructures, and the location of the fultoportulae along the rim(Fig. 1e and f). The wide ribs are typically narrower near the centerof the valve, and widen towards the radius (Fig. 1e and f). The wideribs are most commonly unit structures, but as shown in Fig. 1f,they can be bifurcated and then rejoined at the portula. In betweenthe wide ribs in the region corresponding to the pore field are elon-gated narrow silica ribs (Fig. 1f). Generally 2–4 narrow ribs arepresent between adjacent wide ribs. A cross-sectional view of thevalve (Fig. 1g) shows the relative relief of the fultoportula andthe wide ribs, which become exponentially taller towards the rimof the valve, and have a buttress-like appearance.

Cross-sectioning, AFM, and HF dissolution of the valves revealedtheir internal structure and stages in their formation (Fig. 2). Thethickness of the mature valve measured from SEM cross-sectionsvaries depending on the location of measurement. The central re-gion averaged 169.8 ± 3.5 nm (n = 15), and the rib and pore regionaveraged 128.0 ± 2.6 nm (n = 31). The narrow ribs appeared to beslightly less thick (by 10%) than the wide ribs (Fig. 2b). In all re-gions of the valve, larger silica structures were formed from anamalgamation of smaller irregularly-shaped silica particles of afew nanometers (Fig. 2). In the central region, the amalgamationwas relatively uniform (Fig. 2a), but in the rib and pore regionthe smaller particles were assembled into larger particles that con-stituted the wide and narrow ribs (Fig. 2b and c). Spaces in-be-tween the larger particles constituted the pores. The narrow ribswere somewhat oval in cross-section, but were generally more flat-tened at the proximal surface, and rounded at the distal surface(Fig. 2b and c). This results in a trumpet shape to pores that spanthe cross section (Fig. 2b and c, arrows). Imaging of the ribs duringtheir formation by AFM showed that initially deposited silica wasparticulate, but the surface became smooth with maturation(Fig. 2d). Measurements of roughness in these two regions yieldedRMS values of 2.77 for the initial deposition, and 1.88 for the ma-ture region. These data suggested that the initial structure wasformed from particulate silica in which either the surface of the ini-tially-formed particles became flattened or the spaces in betweenthem became filled in. As measured by AFM, the thickness of thewide ribs at an initial stage of their formation was approximately50 nm, and the narrow ribs were 30 nm (data not shown) – aboutone-third of that measured for the same structures in the maturevalve, which is consistent with thickening.

Application of a synchronized growth procedure (Section 2)enabled examination of valve development using SEM and fluo-rescence microscopy and revealed distinct stages of formation(Fig. 3). Fluorescence microscopy of PDMPO-stained cells(Fig. 3a) was used to identify samples used for further examina-tion by SEM and cells imaged by AFM. Observed at the earlieststage was a fractally-branched structure approximately 4.5–5.5 lm diameter (Fig. 3b and c), which corresponded to the diam-eter of the central region in the mature valve. Precursors to thewide ribs were visible radiating out from the center (Fig. 3band c). At this stage the ends of the ribs did not have portulae.

Page 4: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

Fig. 1. SEM of the valve of Cyclotella cryptica. (a) Overall view of the distal valve surface. Arrow denotes the location of a fultoportula offset from the valve center. (b) View ofthe distal surface, near the outer edge of the valve. Wide ribs with pore fields in between are evident. (c) Central portion of the distal surface. The fultoportula is located withan arrow. (d) Edge-on view of the distal surface, showing the raised pore fields, and location of fultoportulae along the upper rim (arrows). (e) Proximal view of the valve.Wide ribs are predominant and a fultoportula is located by an arrow. (f) Proximal view of the valve rim. Wide ribs are predominant and most contain fultoportulae located onthe valve rim. One wide rib, located by the arrow, is bifurcated. (g) View of proximal surface in a broken valve. This view emphasizes the relief on proximal structures such asthe wide ribs and portulae on the valve edge and in the central region (arrow).

B. Tesson, M. Hildebrand / Journal of Structural Biology 169 (2010) 62–74 65

Page 5: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

Fig. 2. Particulate structures in the valves. (a) SEM of a fractured valve in the central region showing relatively homogeneous nanoparticulate silica. (b) SEM of a fracturedvalve in the rib region. The wide and narrow ribs (w and n, respectively) are denoted with brackets, and the location of a trumpet-shaped pore is shown by an arrow. (c) SEMof a fractured valve in the pore field region. A trumpet-shaped pore is located by an arrow. (b and c) Amalgamation of smaller silica particles into larger structures thatconstitute the ribs. (d) Deflection AFM image of forming wide and narrow ribs. Bracket at lower right denotes the location of initially deposited silica, which is noticeablymore particulate than the more mature silica bracketed at upper left.

Fig. 3. Stages in formation of the valve visualized by SEM and fluorescence microscopy. (a) Fluorescence micrographs of forming valves at different stages visualized byPDMPO incorporation – scale bars are 2 lm. (b and c) Earliest stages, which consists of wide ribs with a fractal organization. (d) Later stages where wide and narrow ribsextend from the central region, and portulae have not formed at the ends of the wide ribs, but initial stages of portula formation are occurring in the central region. (e) Stagewhere pore fields are forming as well as fultoportulae on the wide ribs. (f and g) Fultoportulae formation on the upper valve rim is advanced and pore fields are filling in. Thelower valve rim is being formed. (h) View of proximal surface showing predominance of rib structures (but not pore fields). The lower rim of the valve has not formed, and thesmall ribs are not completely incorporated into the structure at the ends.

66 B. Tesson, M. Hildebrand / Journal of Structural Biology 169 (2010) 62–74

The narrow ribs either branched directly off of the wide ribs, orbranched off of nodes from the wide ribs (Fig. 3c and d). At a laterstage, the wide and narrow ribs extended from the central region,the pore fields were initiated, and the portulae began to form

(Fig. 3e). The wide ribs increased in width radially from the cen-ter, incorporated the fultoportulae at the upper valve rim, andthen became narrower and fused with the lower valve rim(Fig. 3f and g). The edges of the wide ribs contained projections

Page 6: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

B. Tesson, M. Hildebrand / Journal of Structural Biology 169 (2010) 62–74 67

which cross-connected to the narrow ribs (Fig. 3g) – eventuallyforming the distal grid structure. Cross-connections between thenarrow ribs were visible on the distal surface (Fig. 3g), but onlythe ribs were predominant in the proximal view (Fig. 3h). The

Fig. 4. Nanopore structure and arrangement. (a) Deflection AFM of the proximal valve sucircular nanopores in between. (b) Deflection AFM of the central region of the distal valvthe narrow ribs and circular nanopores. (d) Height AFM of HF-treated distal valve surface,form the grid structure.

Fig. 5. Formation and structure of fultoportulae along the valve rim. (a) SEM of initial sforming at the end of bifurcated wide ribs; in the upper central region is a rib that bifurcathat bifurcated closer to the forming fultoportula. (b and c) Deflection AFMs of formingdefined, and (c) is a later stage where rib silica deposition has moved past the location ofportulae showing circularization of the structure. (e and f) SEM of portulae where circchamber is occurring. (g) SEM of completed fultoportulae along the rim of a cross-sectionvisible. The inner and outer projections are not precisely aligned (arrows). (For interpretaversion of this article.)

rim of the valve was formed from the region under the portulae,prior to complete extension of the narrow ribs and pore fields(Fig. 3f). At the final stage the narrow ribs were completely incor-porated into the curvature at the valve rim.

rface highlighting the relatively smooth texture of the wide and narrow ribs, and thee surface. (c) SEM of HF-treated (see Section 2) proximal valve surface, highlightingshowing the lower layer circular nanopore region, and the upper layer deposition to

tage in fultoportula formation. Horseshoe-like structures are visible (green arrows)ted well in advance of the forming fultoportula, and towards the lower right is a ribfultoportulae. (b) is an earlier stage where the initial form of the portula is beingthe portula, and growth of the portula in the z-axis is occurring. (d) SEM of forming

ularization is complete and initial stages of overgrowth of silica to form the outered valve. The projections of the portulae above the proximal and distal surfaces aretion of the references to color in this figure legend, the reader is referred to the web

Page 7: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

68 B. Tesson, M. Hildebrand / Journal of Structural Biology 169 (2010) 62–74

Pores on the proximal and distal valve surfaces had a distinctappearance (Fig. 4). On the proximal surface, they had a smallopening (18 nm) which was circular and conserved in size betweendifferent pores (Fig. 4a). On the distal surface in the central regionof the valve, the nanopores appeared to have more highly raisedrims (Figs. 1c and 4b), but still were relatively circular and consis-tent in size. In the pore fields between the wide ribs, openings werelarger, and irregular in size and shape, with a tendency towardsbeing rectangular (Fig. 1b). Partial etching of valves by hydrofluoricacid (HF) treatment clarified the structure and arrangement of thepores and the valve in general. In the proximal view the circularnanopores were predominant near the valve center (Fig. 4c). Radi-ally towards the periphery of the valve, the narrow ribs were visi-ble, and in-between the ribs were both larger and smaller pores(Fig. 4c). The cross-connections were not visible on the proximal

Fig. 6. SEM of central fultoportulae formation. (a) Early stage where the region of portulaperipheral pores to the central pore are defined. (c) Growth of silica in the central and pesurface showing z-axis extension of the central portion of the portula. (e) View showinsection of a central fultoportula showing the inner chamber and relative extension on t

Fig. 7. Girdle band structures. (a) Smooth portion of the girdle band. On the upper edge iand c) Possible stages in girdle band formation in which a grid-like structure is present (that it consists of assemblies of nanoparticulate silica which are differentiated intohomogeneously organized, and the pore region (towards upper right), in which particle

surface even though the narrow ribs were (Fig. 4c). AFM of theHF-etched distal surface revealed both the grid-like structure andthe nanopores at a lower depth (Fig. 4d). From these images it isapparent that the grid-like structure was a distinct layer made ofparticles of silica raised above the base level of the nanopores(Fig. 4c and d).

3.2. Portulae formation

Fultoportulae on the rim of the valve were formed by separationof the wide ribs into two parts at the margin of the growing valve(Fig. 5). This could involve a branching of the wide ribs either closeto, or well in advance of, the forming portula (Fig. 5a). The initialstage of portula formation was distinguished by the appearanceof a horseshoe-shaped structure (Fig. 5a), which was clearly dis-

formation is defined by a lack of silica deposition (arrow). (b) Later stage where theripheral pores in the z-axis direction has initiated. (d) View from the proximal valveg the lack of extension of the fultoportulae from the distal valve surface. (f) Cross-he distal (upper) and proximal (lower) surface.

s a lip, and the lower edge, which is the interface with the pore region, is serrated. (bb), in what becomes the pore region (c). (d) Deflection AFM of girdle band showing

the smooth region (towards the lower left) in which particles are relativelys are grouped into a rib or grid-like structure.

Page 8: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

B. Tesson, M. Hildebrand / Journal of Structural Biology 169 (2010) 62–74 69

tinct from the wide rib structure. After its initial definition in the x/y plane, the portulae grew in the z-axis direction distally, as can beseen by AFM in Fig. 5b and c. The ribs continued to extend past theportulae (Fig. 5b and c) during z-axis growth. The rim of the portulacircularized (Fig. 5c and d), and eventually became completelyclosed (Fig. 5e). Subsequently, there was an overgrowth from thewide ribs around the portula rim (Fig. 5e and f), which indicatesadditional silica deposition on a higher proximal plane in the z-axisdirection. The overgrowth eventually closed, which generated achamber for the portula that was external to the base layer ofthe valve (see Hildebrand et al., 2006). The portula opening grewslightly more to generate an extended tube (Fig. 5g). The innerand outer projections of the portulae were not precisely aligned(Fig. 5g, arrows).

Fig. 8. Rhodamine phalloidin staining of actin and PDMPO staining of silica associatedcombined actin and silica (green) staining, or actin, silica, and chlorophyll autofluoresceouter rings of actin are visible in the location of the outer edge of the two daughter cell vasilica. (b and b0) Side view of the outer actin ring (o) and the inner filaments (i), which arthe actin and silica are interdigitated. (c and c0) Later stage in valve formation. The outerelation to the growing diameter of the valve. The inner and outer filaments have becombetween actin and silica. (d and d0) Actin remaining associated with the newly formed vathe position and structure of the raised pore field region. (e, e0 , and e0 0) Possible initial staring is distinct from those in (a–c) because it is not the full diameter of the valve. (e) Thoutside of the upper actin ring is a labeled girdle band (GB). (e0 0) Image of the same cell inthe girdle band (which is the extreme outside extent of the cell).

Fig. 9. Rhodamine phalloidin staining of actin and PDMPO staining of silica associated wcells that have not yet separated are visible. Actin is colored purple and newly deposited

The mechanism of the central fultoportula formation had simi-larities and differences compared to the valve rim fultoportulae(Fig. 6). At a very early stage of wide rib formation, the outline ofthe fultoportula structure was defined by an area of inhibition ofsilica deposition (Fig. 6a, arrow). In addition to formation of a largecentral pore, 2–4 smaller peripheral pores surrounding this werepresent (Fig. 6b and c). The peripheral pores were also presenton the valve rim fultoportulae, but were difficult to see becauseof the angle. On the central fultoportula, the tubular structure grewin the z-axis plane on the proximal valve surface (Fig. 6d). Thetubular structure did not extend appreciably on the distal surface(Fig. 6e and f), which is in distinction to the valve rim fultoportulae,and suggests that there is no overgrowth of silica. A cross-section

with valve formation. Sets of images are presented with actin staining (purple),nce (red). Scale bars are 4 lm. (a and a0) Early stage of valve formation. Prominentlves, and inner actin filaments are also visible associated with the initially depositede positioned at the leading edge of silicification. Yellow arrows locate regions wherer actin ring remains in the same position but the inner filaments have expanded ine in close proximity relative to (a). Yellow arrows locate positions of interdigitationlve after daughter cell separation. Yellow arrow locates a loop of actin that mirrorsge in valve (V) formation in which an actin ring surrounds the deposited silica. Thee actin rings in adjacent daughter cells. (e0) Actin and silica, the large loop of silica(e) and (e0) (in 2D, central region) showing the position of the actin rings relative to

ith girdle band formation. In these images, girdle bands in two adjacent daughtersilica is green. Actin forms a ring at the front of silicification. Scale bars are 4 lm.

Page 9: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

70 B. Tesson, M. Hildebrand / Journal of Structural Biology 169 (2010) 62–74

of the fultoportula is visible in Fig. 6f, showing the inner chamberand extension from the proximal and distal valve surfaces.

3.3. Girdle band formation

The detailed structures of girdle bands are rarely imaged (Hilde-brand et al., 2006, 2009b), due to the difficulty of visualizing thesethin structures. It is not clear whether the images presented inFig. 7 represent distinct stages in girdle band formation, but theydo highlight distinct features. The girdle bands in C. cryptica weresimilar in structure to those of T. pseudonana (Hildebrand et al.,2006), and consisted of a relatively featureless portion that over-lapped with adjacent girdle bands, and a more structured part con-taining nanopores. One edge of the featureless portion was smoothand appeared to have a lip, whereas the other edge was serrated(Fig. 7a). A cross-hatched silica grid structure was observed adja-

Fig. 10. SEM of the effects of cytoskeletal inhibitors on valve structure. Treatment witmisalignment of ribs and mispositioning of portulae. Treatment with the microtubuleportulae are mispositioned. Inset in (c) shows that the pore field region blends intocytochalasin D, both distal views. The overgrowth of silica in fultoportula formation is inraised pore field region (Fig. 1d), is not raised after cytochalasin D treatment.

cent to the featureless portion (Fig. 7b) and the ribs were appar-ently precursors to the nanopores (Fig. 7c). AFM revealed thenanoscale features of the girdle bands (Fig. 7d), which consistedof particulate silica even in the relatively featureless portion, andribs that were composed of organized agglomerates of particulatesilica.

3.4. Actin localization

To examine the involvement of the cytoskeleton in cell wall for-mation in C. cryptica, actin was located in relation to newly formedsilica by fluorescence microscopy using rhodamine phalloidin andPDMPO as stains, respectively. Although actin was seen to form adiffuse network inside the cell, it was evident that a substantialaccumulation of actin was located in the proximity of the SDV dur-ing silica polymerization (Fig. 8). In Fig. 8a is a cell at an early stage

h the microtubule inhibitor colchicine, distal (a) and proximal (b) views. Note theinhibitor oryzalin, distal (c) and proximal (d) views. Ribs are also misaligned andthe wide ribs under this treatment. (e and f) Treatment with the actin inhibitorhibited (e), and portions of the valve are not silicified (f). In addition, the normally

Page 10: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

B. Tesson, M. Hildebrand / Journal of Structural Biology 169 (2010) 62–74 71

of valve formation. There are two (one for each daughter cell valve)intensely stained rings of actin located where the periphery of thevalve would be expected (Fig. 8a). In between the two rings wereinner filaments of actin, which were associated with newly depos-ited silica (green fluorescence, Fig. 8a0, right). The inner filamentsand outer ring appeared to be connected (Fig. 8a). Examinationof actin in relation to new silica deposition in another cell con-firmed the connection between the ring and filaments and indi-cated that the inner filaments were located at the leading edge ofsilica deposition, and that some portions of the silica and actinwere interdigitated (Fig. 8b0, yellow arrows). At a later stage ofvalve formation (Fig. 8c) the silica-associated actin was seen tohave expanded (the inner filaments and outer ring are closer to-gether than in Fig. 8a) and yet maintained its peripheral associa-tion with the silica. In addition, strands of actin were seen tointerdigitate between the wide rib/fultoportula portion of the valve(Fig. 8c, yellow arrows). In some cases, the interdigitation was seento persist even after valve completion and daughter cell separation(Fig. 8d, yellow arrows). In several cells we saw what could havebeen the earliest stage in valve formation in which the actin ringwas not positioned at the extreme periphery of the cell (Fig. 8e0 0).It also was in the same plane as the deposited silica, and not atthe curved edge of the completed valve (Fig. 8e0 0). Based on the ini-tial stage structure identified in Fig. 3, this stage could representformation of the central fractally-organized portion of the valve.If so, the data in Fig. 8 would indicate a dynamic movement ofthe actin ring from its initial to final position.

Actin was also visualized associated with forming girdle bands.Cyclotella cryptica starved for silicon can arrest in the G2 phase ofthe cell cycle (unpublished observation), and upon silicon replen-ishment, can complete valve synthesis and separate. In Fig. 9 isan image of cells that are in the process of separation, synthesizingnew girdle bands. In both daughter cells, silica in the girdle bandsis associated with an actin ring (Fig. 9). The actin ring is located atthe edge of the girdle band furthest away from the valve, whichcorresponds to the pore region of the girdle band (Fig. 2).

We attempted without success to localize microtubules in rela-tion to forming silica structures, using both tubulin antibodies andthe microtubule-specific dye Oregon Green paclitaxel (Lillo et al.,2002). As an alternative approach, we examined the effect of twoinhibitors of microtubule polymerization, colchicine and oryzalin,and also the actin inhibitor cytochalasin D, on valve formation(Fig. 10). For all inhibitors, a titration was performed to determinethe concentration at which cell growth was reduced but notstopped, and cells from that treatment were isolated and purifiedfor SEM. Although these inhibitors could have multiple cellular ef-fects depending on the specific process an individual cell wasundergoing, several consistent trends on alteration of valve struc-ture were observed. Both microtubule inhibitors (which have sim-ilar modes of action) affected the positioning (but not necessarilythe structure) of the fultoportulae, which were frequently locatedshort of their normal location at the valve edge (Fig. 10b–d). Themicrotubule inhibitors also disrupted both the width and the radialorientation of the wide ribs, leading to a less organized and less lin-ear arrangement (Fig. 10a–d). Close examination showed that inregions where the wide ribs were diminished, the pore fields ex-tended into them (Fig. 10c, inset), and in general there was an in-crease in the numbers of thin ribs. The actin inhibitor cytochalasinD did not affect the positioning of the fultoportulae, but did affectthe ‘‘overgrowth” of silica involved in forming the outer chamberof the fultoportula (forming an elongated slit rather than a circularopening), and had a slight effect on the linearity of the wide ribs(Fig. 10e and f). In some instances, portions of the valve structurewere entirely compromised (Fig. 10f). The pore field regions werealso affected, and frequently were not well formed or were notraised as high above the valve surface as usual (Fig. 10e).

4. Discussion

The present study on cell wall structure formation in C. crypticarevealed differences in structure formation compared with a previ-ous analysis on T. pseudonana (Hildebrand et al., 2006, 2007). TheCyclotella cryptica valve has distinct a three-dimensional compo-nent, with pore fields raised above the plane of the rest of the valve(Fig. 1d) and has two distinct forms of ribs, both wide and narrow,whereas T. pseudonana has a flat valve and only consistently-sizednarrow ribs (Hildebrand et al., 2006). The differences in structurerelate to differences in structure formation processes in thesetwo species.

Valve formation in C. cryptica demonstrates a fractal to path-dependent transition comparing the central region with the widerib and pore field region (Fig. 3). At a fundamental level this sug-gests a transition from a relatively static self-organizing processin the central region, to a dynamic directional growth process toform the wide ribs and pore fields (Fig. 3). The change in positionof actin in concert with the growing silica polymerization front(Fig. 8) suggests that movements of actin are involved in the direc-tional dynamic growth. In this scenario, initial silica deposition isconfined within a space delineated by an actin ring (Fig. 8e) whichcorresponds to the primary silicification site (PSS) identified inother diatom species (Pickett-Heaps et al., 1990). There are appar-ently two stages of expansion of actin after the initial silica deposi-tion. The first involves the expansion of the ring to its final positionat the edge of the valve (compare Fig. 8e and 8a), and this expansiondoes not include the substantial deposition of silica. Recent work inT. pseudonana also indicated that the SDV was fully expanded priorto complete silica deposition, and showed an association with ac-tin-sized filaments (Hildebrand et al., 2009a). The reason for thetwo-step positioning of the actin ring in C. cryptica is not clear, how-ever we can speculate that it may have something to do with en-abling formation of a flat-on-the-top (initial stage) but curved-on-the-edge (later stage) valve structure where the rigid flat structureprovides an anchor point to enable curvature at the rim. Detailedinvestigation of other species that make distinctly-shaped valvesis needed to provide further insight into this. The second processof actin expansion is seen in the movements of the actin filamentsclosely associated with the silica (Fig. 8a–c). Association of filamen-tous actin with the leading edge of silicification has been shown byTEM in a variety of diatom species (Pickett-Heaps et al., 1990; vande Meene and Pickett-Heaps, 2002, 2004), indicating a conservedand essential functional role. The filamentous actin may play animportant role in shaping or maintaining the integrity of the SDVmembrane, as will be discussed in more detail below.

Cross-sectioning and AFM revealed that all structures in C. cryp-tica are formed from nanoparticulate silica (Fig. 2), and theseassemble into larger particles which are organized on an evenhigher level to form mesoscale structure. Nanoparticulate silicahas been observed previously in diatom valve cross-sections andon surfaces (Crawford et al., 2001; Hildebrand et al., 2006; Losicet al., 2007; Noll et al., 2002). In some cases, these were shownto have species-specific average particle diameters (Crawfordet al., 2001), which could result from the properties of silica poly-merization determinants such as silaffins, LCPAs, and silacidins(Kröger et al., 2000, 2002; Poulsen and Kröger, 2004; Wenzlet al., 2008) unique to the species. Particle formation and associa-tion is a thermodynamically-favored aspect of silica polymeriza-tion chemistry (Iler, 1979). In diatoms, particle formation iscatalytically promoted by the interaction between silaffins, LCPAs,and silacidins, however the silica particles do not spontaneouslyself-organize into a higher order (mesoscale) structure, althoughtheir assembly can reflect nanoscale textures seen in diatom frus-tules (Kröger et al., 2000, 2002; Poulsen and Kröger, 2004; Wenzlet al., 2008). Thus, in diatoms other organizational determinants

Page 11: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

Fig. 11. Model for fultoportula formation. (a) Detailed view of proximal portion of avalve rim fultoportula showing the large central pore flanked by three accessorypores. (b) Schematic of how proposed organic complexes positioning the centraland accessory pores control the molding of forming silica around them.

72 B. Tesson, M. Hildebrand / Journal of Structural Biology 169 (2010) 62–74

must be involved in the formation of mesoscale structure (Davisand Hildebrand, 2007; Pickett-Heaps et al., 1990; Robinson andSullivan, 1987). The mesoscale assembly pattern of particles asseen here (Fig. 2) is consistent with either confinement of silicapolymerization within a defined space and/or organization of silicapolymerization determinants on a template with numerous nucle-ating sites. Recent work in T. pseudonana is consistent with theconcept of confinement in the formation of valve structure, andtemplating with the girdle bands (Hildebrand et al., 2009a).

An interesting observation is the smoothing of the surface ofstructures as they mature (Fig. 2). The cross-sections (Fig. 2b and c)reveal that part of the smooth surface morphology, especially onthe proximal valve face, is intrinsically generated during silica parti-cle formation, perhaps due to close oppression of the polymerizingsilica with the SDV membrane (Crawford et al., 2001; Hildebrandand Wetherbee, 2003; Pickett-Heaps et al., 1990). On the distal sur-face, this is not the case (Fig. 2b). This is consistent with a closer asso-ciation of one side of the membrane with the silica, and alsoadditional deposition after initial particle formation on the otherside. It is well established in other diatom species that at initial stagesof deposition, silica is most closely associated with one side of theSDV, and that the SDV expands in the z-axis only in one direction to-wards the other side (Hildebrand et al., 2006; Schmid and Schulz,1979; Schmid and Volcani, 1983). In some cases, the morphologyof silica formed during z-axis expansion is dramatically differentfrom the morphology of initially deposited silica (Hildebrand et al.,2006). In C. cryptica the z-axis expansion is apparently minimal.

On the proximal valve surface, circular nanopores (18 nm diam-eter) are visible in-between the ribs in the pore field region(Fig. 4a). On the distal surface, circular pores with similar charac-teristics are visible in the central region of the valve (Fig. 4b), butin the raised pore field region pores tend to be larger in diameterand elongated in the radial direction – some having a rectangularshape (Fig. 1b). The pore field regions (and wide and narrow ribs)constitute a distinct silica layer from the nanoporous region as canbe seen in Fig. 4c and d. The distinct layer is not clearly evident inthe cross-sections of Fig. 2a–c, which could be due to (1) the pre-cise location of the viewed image, (2) disruption of the additionallayer during cleavage, or (3) complete integration of the additionallayer in the fully mature structure. From the proximal view, thenarrow ribs are seen as linear structures aligned radially on thevalve (Figs. 1f and 4c). On the distal surface the radially-orientedribs are still visible, but much more noticeable are cross-connec-tions between the ribs (Fig. 1b). The lack of cross-connections vis-ible from the proximal side is consistent with a distinct silica layerobscuring the view, and the height difference between the nanop-ores and ribbed structures visible by AFM in Fig. 4d, directly showsthe layers. The cross-connected rib structure is formed at an earlystage during the expansion of the valve from the central region(Fig. 3g), and the nanopore layer is apparently formed later.

SEM and AFM provide a detailed look at fultoportulae formation(Figs. 5 and 6). The horseshoe structure of the initial form of the por-tula is well defined, and has the appearance of a pre-structured fea-ture around which silica is deposited as the wide ribs are forming(Fig. 5a–d). A reasonable hypothesis is that the fultoportulae areformed from an organic complex that is pre-positioned at specificlocations where the wide ribs will propagate and form prior to silicadeposition around the complex (Fig. 11). We interpret the differ-ences in diameter of the forming portulae in Fig. 5b and c as beingdue to destruction of an organic complex at the earlier stage inFig. 5b resulting from acid extraction. Recently, Davis and Hilde-brand (2007) have suggested a model where silica polymerizationdeterminants are anchor to the silicalemma via interaction withtransmembrane proteins and the cytoskeleton, the same mecha-nism could be involved in the pre-positioning of the portulae com-plex. The results of the inhibitor experiments (Fig. 10), suggest that

microtubules are involved in this positioning. To our knowledge, noTEM cross-sectional imaging of organics associated with the portu-lae during their formation have been published (although images ofthe complete portula structure have been – Herth, 1979, 1979),however other organic complexes have been seen associated withother silica structures projecting from the valve plane (Li and Vol-cani, 1985; Pickett-Heaps et al., 1990), suggesting that a proposedfultoportula complex is reasonable. The growing fultoportula formsboth distal and proximal extensions from the valve surface, and inthe case of the distal extensions, involve an overgrowth of silicafrom additional deposition on the wide ribs to form a projectingchamber (Fig. 5e–g, and Hildebrand et al., 2006). Such overgrowthis minimal in the central fultoportula (Fig. 6f), perhaps becausegrowth is not path-dependent in that region.

We cannot rigorously establish whether the data in Fig. 7 repre-sent distinct stages in girdle band formation. However, the locali-zation of actin in Fig. 9 and the previous localization of actinalways at the leading edge of silicification (Fig. 8 and Pickett-Heapset al., 1990; van de Meene and Pickett-Heaps, 2002, 2004) is con-sistent with the relatively unstructured portion of the girdle bandsbeing synthesized first, followed by the nanopore region. Theunstructured portion is the region of girdle band that ‘‘underlaps”with the adjacent girdle band, and it makes logical sense that thiswould be synthesized first to maintain structural integrity. RecentAFM data (Hildebrand et al., 2009b) coupled with these images(Fig. 7) indicate that girdle bands exhibit similar complexity intheir structure and formation processes as some valve structures.

Our results indicate a significant role for actin in the formationof both valves and girdle bands (Figs. 9 and 10). In both structures,actin is located at the leading edge of silicification (Figs. 9 and 10)suggesting that it is a critical determinant of silica formation. It hasbeen proposed that actin filaments are involved in the extensionand morphogenesis of the SDV through association with the grow-ing edge of the silicalemma (Pickett-Heaps et al., 1990). Detailedobservations of actin localization by fluorescence microscopy havebeen done in the centric diatoms Proboscia alata and Rhizoseleniasetigera (van de Meene and Pickett-Heaps, 2002, 2004). These spe-cies have elongated and conical valve shapes, in contrast to the rel-atively flat-on-the-surface but curved-at-the-perimeter valve of C.cryptica and many other centric species. All three species have sub-stantial actin rings, and in P. alata and R. setigera the rings have

Page 12: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

Fig. 12. Interdigitation of structures in valves of adjacent daughter cells. (a) SEM ofregion where daughter cells are forming. Arrows locate the position of fultoportulaein adjacent valves. (b) Fluorescence image of PDMPO-stained silica, showing thealternating position of fultoportulae in adjacent valves. Scale bar is 2 lm.

B. Tesson, M. Hildebrand / Journal of Structural Biology 169 (2010) 62–74 73

been shown to be dynamically moved along with the silicificationfront and expand in diameter with the increase in diameter of thevalve during its formation (van de Meene and Pickett-Heaps, 2002,2004). In C. cryptica there appear to be two positions for the actinring; the first defines the location of initial deposition, and the sec-ond the edge of the completed valve (Fig. 8). At the second posi-tion, the ring is not associated with silica (Fig. 8). The additionalactin network associated with the silicification front in C. cryptica(Fig. 8) apparently plays a similar dynamic role as described for ac-tin in the other species at the leading edge of silicification.

There are different possible functional roles that actin couldplay in the silicification process. One is as a stabilizer of the SDVmembrane to prevent osmotically-induced swelling and breakage.Such a role has been documented in the maintenance of the Golgiapparatus in its stacked form – actin networks maintain the flat-tened membrane shape and prevent swelling resulting from accu-mulation of high concentrations of osmolytes (proteins andcarbohydrates) in the Golgi lumen (di Campli et al., 1999; Egeaet al., 2006). Osmotic stress should also be high in the SDV consid-ering the accumulation of solid silica material inside the vesicle,which should induce swelling. One possibility is that actin stabi-lizes the SDV membrane until silica is deposited, at which pointthe membrane becomes anchored to the mineral and actin is nolonger needed. A recently described close association of the SDVmembrane only where silica was deposited is consistent with this(Hildebrand et al., 2009a). The effect of cytochalasin D on the over-growth of silica during fultoportula formation but not the position-ing of the ribs or portulae (Fig. 10e and f), suggests that actin mayplay a minimal role in the initial deposition of ribs, but a more sub-stantial role in the subsequent expansion of the additional silicalayer (Fig. 4). A second possible role for actin is in membrane shap-ing, in which actin is known to be involved in other systems (Egeaet al., 2006; Lanzetti, 2007). Considering the interdigitation be-tween actin and the silica adjacent to the raised pore field regions(Fig. 8), the hemispherical shape of actin observed in Fig. 8d, whichhas the same curvature as the silica in the pore field region, and theflattening of the pore field region in the presence of the actin inhib-itor (Fig. 10e), actin could be involved in shaping the SDV mem-brane prior to deposition in this region. However, an alternateexplanation is that actin is only playing a membrane stabilizingrole, and conforms to the membrane shape derived from otherforces. A third possibility relates to previous studies (Pickett-Heapset al., 1990) and our results (Fig. 8) that indicate an involvement ofactin in the microscale overall shaping of the forming silica struc-ture. Models have been presented (Davis and Hildebrand, 2007;Robinson and Sullivan, 1987) whereby cytoskeletal elements orga-nize silica polymerization determinants via interaction with pro-teins spanning the silicalemma, similar to the actin/spectrin/integrin system (Branton et al., 1981), thus actin could be involvedin organizing silica polymerization determinants in the SDV, andplay a role in structure formation at the mesoscale.

The mechanisms involved in formation of the raised pore fieldsrelate to how diatoms may form particular three-dimensionalstructures. The raised regions must require extra addition of SDVmembrane in their locations. In C. cryptica, the wide rib, fultoport-ulae, and valve rim (which may form an inflexible framework) areformed prior to completion of the raise pore fields (Fig. 3). As thepore field regions are developed, SDV membrane componentscould be added in excess in these regions, which could result in aneed to bend the structure above the plane of wide rib deposition.Why would it bend up instead of down? Because of the direction-ality of membrane trafficking, membrane material must be addedeither on the proximal side of the SDV or at the leading edge ofsilicification. If the proximal membrane is relatively ‘‘fixed” bypositioning of silica polymerization determinants or polymerizedsilica, then excess ‘‘free” membrane material would accumulate

on the distal surface, enabling the structure to bend that way. An-other contribution could be ‘‘interactive division” (Mann, 1984), inwhich adjacent daughter cell valves are closely appressed, andnon-planar structures result in protrusions in one valve that formdepressions in the other. We observed such an alignment betweenadjacent daughter cell valves in C. cryptica (Fig. 12). Interactivedivision probably only lowers the energy requirements for forma-tion of such structures; however, and does not dictate them be-cause there is no previously fixed structure in the formation ofinitial cells after sexual reproduction.

Our data (Figs. 8–10), coupled with previous observations (Pick-ett-Heaps et al., 1990), suggest several underlying concepts regard-ing valve formation in C. cryptica, and perhaps in general for centricdiatoms with circular valves. Microtubules have been universallyassociated with rib structures in diatoms (Pickett-Heaps et al.,1990), and our data are consistent with microtubules definingthe location of the wide ribs and positioning the fultoportulae(Fig. 10). Previous work on P. alata suggested that microtubules

Page 13: Journal of Structural Biology - Research Profilesscrippsscholars.ucsd.edu/.../files/...j_struc_biol.pdf · at the single cell level. The cell biology of diatom silicification has

74 B. Tesson, M. Hildebrand / Journal of Structural Biology 169 (2010) 62–74

were involved in the radial elongation of the valve (van de Meeneand Pickett-Heaps, 2002). Actin is universally associated with thesilica polymerization front (Pickett-Heaps et al., 1990; van deMeene and Pickett-Heaps, 2002, 2004, and Figs. 8 and 9), andalthough there may be multiple reasons for this (see previous par-agraph), one consistent observation is that actin defines the diam-eter of the circular valve, and expansion of actin results inexpansion of the silicification front. It is possible that actin inter-acts with microtubules to either help define the shape of the SDVin a static sense or enable dynamic movement of the silicificationfront during expansion. In the presence of microtubule or actininhibitors (Fig. 10), patterning was affected but silicification stilloccurred. This indicates that additional components besides thecytoskeleton are involved in mesoscale structure formation, butsuggests that an interaction between these components and thecytoskeleton are essential for correct pattern formation.

Acknowledgments

This research was supported by Air Force Office of Scientific Re-search Multidisciplinary University Research Initiative GrantRF00965521. We are grateful to Farooq Azam and Francesca Mal-fatti for use and assistance with AFM. AFM was funded by a Gordonand Betty Moore Foundation award to Farooq Azam.

References

Addadi, L., Weiner, S., 1992. Control and design principles in biologicalmineralization. Angew. Chem. Int. Ed. Engl. 31, 153–169.

Apt, K.E., Kroth-Pancic, P.G., Grossman, A.R., 1996. Stable nuclear transformation ofthe diatom Phaeodactylum tricornutum. Mol. Gen. Genet. 252, 572–579.

Armbrust, E.V., Berges, J.A., Bowler, C., Green, B.R., Martinez, D., Putnam, N.H., Zhou,S.G., Allen, A.E., Apt, K.E., Bechner, M., Brzezinski, M.A., Chaal, B.K., Chiovitti, A.,Davis, A.K., Demarest, M.S., Detter, J.C., Glavina, T., Goodstein, D., Hadi, M.Z.,Hellsten, U., Hildebrand, M., Jenkins, B.D., Jurka, J., Kapitonov, V.V., Kroger, N.,Lau, W.W.Y., Lane, T.W., Larimer, F.W., Lippmeier, J.C., Lucas, S., Medina, M.,Montsant, A., Obornik, M., Parker, M.S., Palenik, B., Pazour, G.J., Richardson, P.M.,Rynearson, T.A., Saito, M.A., Schwartz, D.C., Thamatrakoln, K., Valentin, K., Vardi,A., Wilkerson, F.P., Rokhsar, D.S., 2004. The genome of the diatom Thalassiosirapseudonana: ecology, evolution, and metabolism. Science 306, 79–86.

Bowler, C., Allen, A.E., Badger, J.H., Grimwood, J., Jabbari, K., Kuo, A., Maheswari, U.,Martens, C., Maumus, F., Otillar, R.P., 2008. The Phaeodactylum genome revealsthe evolutionary history of diatom genomes. Nature 456, 239–244.

Branton, D., Cohen, C.M., Tyler, J., 1981. Interaction of cytoskeletal proteins on thehuman-erythrocyte membrane. Cell 24, 24–32.

Cohn, S.A., Nash, J., Pickett-Heaps, J.D., 1989. The effect of drugs on diatom valvemorphogenesis. Protoplasma 149, 130–143.

Crawford, S.A., Higgins, M.J., Mulvaney, P., Wetherbee, R., 2001. Nanostructure ofthe diatom frustule as revealed by atomic force and scanning electronmicroscopy. J. Phycol. 37, 543–554.

Davis, A.K., Hildebrand, M., 2007. Molecular processes of biosilicification in diatoms.Met. Ions Life Sci. 4, 255–294.

di Campli, A., Valderrama, F., Babia, T., De Matteis, M.A., Luini, A., Egea, G., 1999.Morphological changes in the Golgi complex correlate with actin cytoskeletonrearrangements. Cell Motil. Cytoskeleton 43, 334–348.

Drum, R.W., Pankratz, H.S., 1964. Post mitotic fine structure of Gomphonemaparvulum. J. Ultrastruct. Res. 10, 217–223.

Dunahay, T.G., Jarvis, E.E., Roessler, P.G., 1995. Genetic transformation of thediatoms Cyclotella cryptica and Navicula saprophila. J. Phycol. 31, 1004–1012.

Egea, G., Lazaro-Dieguez, F., Vilella, M., 2006. Actin dynamics at the Golgi complexin mammalian cells. Curr. Opin. Cell Biol. 18, 168–178.

Herth, W., 1979a. A special chitin-fibril-synthesizing apparatus in the centricdiatom Cyclotella. Naturwissenschaften 65, 260–261.

Herth, W., 1979b. The site of b-chitin fibril formation in centric diatoms. II. Thechitin forming cytoplasmic structures. J. Ultrastruct. Res. 68, 16–27.

Hildebrand, M., 2008. Diatoms, biomineralization processes and genomics. Chem.Rev. 108, 4855–4874.

Hildebrand, M., Wetherbee, R., 2003. Components and control of silicification indiatoms. In: Müller, W.E.G. (Ed.), Progress in Molecular and Subcellular Biology,Silicon Biomineralization. Springer-Verlag, Berlin, pp. 11–57.

Hildebrand, M., Frigeri, L.G., Davis, A.K., 2007. Synchronized growth of Thalassiosirapseudonana (Bacillariophyceae) provides novel insights into cell-wall synthesisprocesses in relation to the cell cycle. J. Phycol. 43, 730–740.

Hildebrand, M., Kim, S., Shi, D., Scott, K., Subramaniam, S., 2009a. 3D imaging ofdiatoms with ion-abrasion scanning electron microscopy. J. Struct. Biol. 166,316–328.

Hildebrand, M., Holton, G., Joy, D.C., Doktycz, M.J., Allison, D.P., 2009b. Diverse andconserved nano- and mesoscale structures of diatom silica revealed by atomicforce microscopy. J. Microsc. 235, 172–187.

Hildebrand, M., York, E., Kelz, J.I., Davis, A.K., Frigeri, L.G., Allison, D.P., Doktycz, M.J.,2006. Nanoscale control of silica morphology and three-dimensional structureduring diatom cell wall formation. J. Mater. Res. 21, 2689–2698.

Horcas, I., Fernandez, R., Gomez-Rodriguez, J.M., Colchero, J., Gomez-Herrero, J.,Baro, A.M., 2007. WSXM: a software for scanning probe microscopy and a toolfor nanotechnology. Rev. Sci. Instrum. 78, 013705.

Iler, R.K., 1979. The Chemistry of Silica: Solubility, Polymerization, Colloid andSurface Properties, and Biochemistry. Wiley-Interscience, New York. 866 pp..

Kröger, N., Deutzmann, R., Sumper, M., 1999. Polycationic peptides from diatombiosilica that direct silica nanosphere formation. Science 286, 1129–1132.

Kröger, N., Deutzmann, R., Sumper, M., 2001. Silica-precipitating peptides fromdiatoms – the chemical structure of silaffin-1A from Cylindrotheca fusiformis. J.Biol. Chem. 276, 26066–26070.

Kröger, N., Deutzmann, R., Bergsdorf, C., Sumper, M., 2000. Species-specificpolyamines from diatoms control silica morphology. Proc. Natl. Acad. Sci. USA97, 14133–14138.

Kröger, N., Lorenz, S., Brunner, E., Sumper, M., 2002. Self-assembly of highlyphosphorylated silaffins and their function in biosilica morphogenesis. Science298, 584–586.

Kröger, N., Poulsen, N., 2008. Diatoms – from cell wall biogenesis tonanotechnology. Annu. Rev. Genet. 42, 83–107.

Lanzetti, L., 2007. Actin in membrane trafficking. Curr. Opin. Cell Biol. 19, 453–458.Lenoci, L., Camp, P.J., 2008. Diatom structures templated by phase-separated fluids.

Langmuir 24, 217–223.Li, C.-W., Volcani, B.E., 1985. Studies on the biochemistry and fine structure of silica

shell formation in diatoms. VIII. Morphogenesis of the cell wall in a centricdiatom, Ditylum brightwellii. Protoplasma 124, 10–29.

Lillo, M.P., Canadas, O., Dale, R.E., Acuna, A.U., 2002. Location and properties of thetaxol binding center in microtubules: a picosecond laser study with fluorescenttaxoids. Biochemistry 41, 12436–12449.

Losic, D., Mitchell, J.G., Voelcker, N.H., 2009. Diatomaceous lessons innanotechnology and advanced materials. Adv. Mater. 21, 1–12.

Losic, D., Pillar, R.J., Dilger, T., Mitchell, J.G., Voelcker, N.H., 2007. Atomic forcemicroscopy (AFM) characterization of the porous silica nanostructure of the twocentric diatoms. J. Porous Mater. 14, 61–69.

Mann, D.G., 1984. An ontogenetic approach to diatom systematics. In: Mann, D.G.(Ed.), Proceedings of the Seventh International Diatom Symposium. O. Koeltz,Koenigstein, pp. 113–141.

Noll, F., Sumper, M., Hampp, N., 2002. Nanostructure of diatom silica surfaces and ofbiomimetic analogues. NanoLetters 2, 91–95.

Parkinson, J., Brechet, Y., Gordon, R., 1999. Centric diatom morphogenesis: a modelbased on a DLA algorithm investigating the potential role of microtubules.Biochim. Biophys. Acta 1452, 89–102.

Pickett-Heaps, J., Schmid, A.-M.M., Edgar, L.A., 1990. The cell biology of diatom valveformation. In: Round, F.E., Chapman, D.J. (Eds.), Progress in PhycologicalResearch. Biopress Ltd., Bristol, pp. 1–168.

Poulsen, N., Kröger, N., 2004. Silica morphogenesis by alternative processing ofsilaffins in the diatom Thalassiosira pseudonana. J. Biol. Chem. 279, 42993–42999.

Robinson, D.H., Sullivan, C.W., 1987. How do diatoms make silicon biominerals?TIBS 12, 151–154.

Sandhage, K.H., Allan, S.M., Dickerson, M.B., Gaddis, C.S., Shian, S., Weatherspoon,M.R., Cai, Y., Ahmad, G., Haluska, M.S., Snyder, R.L., Unocic, R.R., Zalar, F.M.,Zhang, Y., Rapp, R.A., Hildebrand, M., Palenik, B.P., 2005. Merging biological self-assembly with synthetic chemical tailoring: the potential for 3-D genetically-engineered micro/nano-devices (3-D GEMS). Int. J. Appl. Ceram. Tech. 2, 317–326.

Schmid, A.-M.M., Schulz, D., 1979. Wall morphogenesis in diatoms: deposition ofsilica by cytoplasmic vesicles. Protoplasma 100, 267–288.

Schmid, A.-M.M., Volcani, B.E., 1983. Wall morphogenesis in Coscinodiscus wailesii. I.Valve morphology and development of its architecture. J. Phycol. 19, 387–402.

Schmid, A.-M.M., 1984. Wall morphogenesis in Thalassiosira eccentrica: comparisonof auxospore formation and the effect of MT-inhibitors. In: Mann, D.G. (Ed.),Proceedings of the Seventh International Diatom Symposium. O. Koeltz,Koenigstein, pp. 47–70.

Shimizu, K., Del Amo, Y., Brzezinski, M., Stucky, G.D., Morse, D.E., 2001. A novelfluorescent silica tracer for biological silicification studies. Chem. Biol. 8, 1051–1060.

Sumper, M., 2002. A phase separation model for the nanopatterning of diatombiosilica. Science 295, 2430–2433.

Sumper, M., Kröger, N., 2004. Silica formation in diatoms: the function of long-chainpolyamines and silaffins. J. Mater. Chem. 14, 2059–2065.

van de Meene, A.M.L., Pickett-Heaps, J.D., 2002. Valve morphogenesis in the centricdiatom Proboscia alata Sundstrom. J. Phycol. 38, 351–363.

van de Meene, A.M.L., Pickett-Heaps, J.D., 2004. Valve morphogenesis in the centricdiatom Rhizosolenia setigera (Bacillariophyceae, Centrales) and its taxonomicimplications. Eur. J. Phycol. 39, 93–104.

Wenzl, S., Hett, R., Richthammer, P., Sumper, M., 2008. Silacidins: highly acidicphosphopeptides from diatom shells assist in silica precipitation in vitro.Angew. Chem. 47, 1729–1732.